Next Article in Journal
Transarterial Chemoembolization with BioPearls for the Treatment of Hepatocellular Carcinoma: A Preliminary Experience
Next Article in Special Issue
Antidiabetic GLP-1 Receptor Agonists Have Neuroprotective Properties in Experimental Animal Models of Alzheimer’s Disease
Previous Article in Journal
The Regulatory Role of CTCF in IL6 Gene Transcription Assessed in Breast Cancer Cell Lines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Small-Molecule Inhibitors of Amyloid Beta: Insights from Molecular Dynamics—Part A: Endogenous Compounds and Repurposed Drugs

by
Mariyana Atanasova
Faculty of Pharmacy, Medical University of Sofia, 1000 Sofia, Bulgaria
Pharmaceuticals 2025, 18(3), 306; https://doi.org/10.3390/ph18030306
Submission received: 10 February 2025 / Revised: 19 February 2025 / Accepted: 21 February 2025 / Published: 23 February 2025

Abstract

:
The amyloid hypothesis is the predominant model of Alzheimer’s disease (AD) pathogenesis, suggesting that amyloid beta (Aβ) peptide is the primary driver of neurotoxicity and a cascade of pathological events in the central nervous system. Aβ aggregation into oligomers and deposits triggers various processes, such as vascular damage, inflammation-induced astrocyte and microglia activation, disrupted neuronal ionic homeostasis, oxidative stress, abnormal kinase and phosphatase activity, tau phosphorylation, neurofibrillary tangle formation, cognitive dysfunction, synaptic loss, cell death, and, ultimately, dementia. Molecular dynamics (MD) is a powerful structure-based drug design (SBDD) approach that aids in understanding the properties, functions, and mechanisms of action or inhibition of biomolecules. As the only method capable of simulating atomic-level internal motions, MD provides unique insights that cannot be obtained through other techniques. Integrating experimental data with MD simulations allows for a more comprehensive understanding of biological processes and molecular interactions. This review summarizes and evaluates MD studies from the past decade on small molecules, including endogenous compounds and repurposed drugs, that inhibit amyloid beta. Furthermore, it outlines key considerations for future MD simulations of amyloid inhibitors, offering a potential framework for studies aimed at elucidating the mechanisms of amyloid beta inhibition by small molecules.

1. Introduction

Dementia is a term associated with memory loss, difficulties in thinking, learning, judgment, and social abilities. It ranks seventh among the leading causes of death across all age groups and fourth among elderly individuals over 70 years old [1]. According to the WHO, as of 2023, more than 55 million people worldwide are living with dementia, with this number increasing by 10 million annually [2]. Alzheimer’s disease (AD) is the most prevalent form of dementia, accounting for 60–70% of cases globally [2]. Additionally, AD is the most common form of amyloidosis, a group of diseases characterized by the abnormal deposition of fibrillar proteins [3]. The pathological hallmarks of AD include senile plaques—extracellular deposits of amyloid β (Aβ) peptide—and intracellular neurofibrillary tangles (NFTs) composed of hyperphosphorylated tau (τ) protein. Moreover, the complexity of the disease involves oxidative stress, which triggers inflammation and leads to Aβ- and tau-induced neurodegeneration [4,5]. Other key factors include the low levels of the neurotransmitter acetylcholine (ACh) in synaptic junctions, resulting in cognitive dysfunction, as well as synaptic and neuronal loss [6,7,8]. AD primarily affects cortical and hippocampal neurons in the human brain. One current therapeutic approach is based on the cholinergic hypothesis, which involves inhibiting the enzyme acetylcholinesterase (AChE) to increase ACh levels, thereby improving cholinergic function and cognition [9,10,11]. Another approved treatment targets overstimulated N-methyl-D-aspartate (NMDA) receptors through the antagonist memantine, which reduces hyperexcitability and prevents neuronal death [4]. However, these treatments are only symptomatic and do not prevent disease progression. As a result, alternative strategies for discovering new anti-AD agents—such as those based on the amyloid cascade and tau hypotheses—are currently the focus of research [12,13,14,15,16]. Despite extensive studies, the pathogenesis of the disease remains to be fully elucidated.
This review is organized into three main sections. The first section examines the amyloid hypothesis along with the structural characteristics and forms of peptide monomers, dimers, oligomers, and their stabilizing interactions. The second section focuses on molecular dynamics simulations, addressing challenges related to timescale limitations and sampling deficiencies. Finally, the third section summarizes and discusses molecular dynamics studies from the past decade on small molecules, specifically endogenous compounds and repurposed drugs, that inhibit amyloid beta.

2. Amyloid Hypothesis

The amyloid hypothesis currently serves as the leading model of AD pathogenesis [12]. According to this hypothesis, Aβ peptide is the central factor that induces neurotoxicity and triggers of pathological events within the central nervous system (CNS). The primary pathological change observed in the AD brain is the aggregation of Aβ into oligomers and deposits, which initiate processes such as vascular damage, activation of astrocytes and microglia due to inflammation, and disruption of neuronal ionic homeostasis and oxidative stress. These changes lead to altered kinase or phosphatase activities, tau phosphorylation, neurofibrillary tangle (NFT) formation, cognitive decline, synaptic loss, cell death, and dementia [12,17]. Additionally, it has been observed that toxic phosphorylated tau forms exacerbate Aβ toxicity through a feedback loop [18]. The amyloid hypothesis is further supported by the mutations in the APP (amyloid precursor protein) gene, as well as in the PSEN1 and PSEN2 genes, which encode the catalytic subunits of γ-secretase [19,20]. These mutations are associated with dominant, early-onset familial AD (FAD). Furthermore, non-dominant, late-onset sporadic AD (SAD) is linked to genetic factors such as the APOE ε4 (apolipoprotein E ε4) allele [21,22,23], as well as issues with Aβ clearance [24] and increased BACE-11 (β—secretase 1, β-site amyloid precursor protein cleaving enzyme 1) activity [25] (Figure 1). The Aβ peptide is produced through the cleavage of the amyloid-β precursor protein, APP. There are three major forms of APP produced by the APP gene, which vary in length and location. These include APP695, consisting of 695 residues and primarily produced in neurons; APP751, composed of 751 amino acids (aa) and expressed in endothelial cells; and APP770, made up of 770 aa and found in platelets [26]. APP is a type I transmembrane glycoprotein composed of three domains: (i) an extracellular glycosylated N-terminal ectodomain (residues 18–624), (ii) a single hydrophobic transmembrane domain (TMD) (residues 625–648), and (iii) a short intracellular domain (AICD) (residues 649–695) [27,28,29,30]. APP undergoes two proteolytic cleavage pathways (Figure 1). In the non-amyloidogenic pathway, cleavage occurs in two steps, first by α-secretase and then by γ-secretase. The α-secretase cleavage results in the release of the soluble ectodomain sAPPα into the extracellular space. The remaining C-terminal fragment, CTFα (83 amino acids), remains embedded in the membrane, where γ-secretase cleaves it, producing the fragment p3 and the APP intracellular domain (AICD), which are released into the extracellular space and cytoplasm, respectively. In the amyloidogenic pathway, β-secretase cleaves APP to generate sAPPβ, which is released into the extracellular space, while the C99 (CTFβ) fragment remains anchored in the membrane. Next, γ-secretase cleaves C99 into Aβ, which is released from the cell, while the AICD is released into the cytosol (Figure 1) [23,31,32]. The Aβ molecule is a 4 kDa polypeptide consisting of 38–53 amino acids [33]. Among the different Aβ isoforms, Aβ1–42 is considered the most neurotoxic [34,35], making it the primary focus of this review. It undergoes sequential self-aggregation, forming various structures such as soluble oligomers, insoluble protofibrils, and fibrils through a primary nucleation mechanism. Three potential pathways for the self-aggregation of Aβ have been proposed, based on the initial secondary structure of Aβ1–42 (Figure 1) [36,37]. The toxic amyloid pathway begins with the primary nucleation of U-shaped β-strand monomers into nuclei during the lag phase. A key early step in amyloidogenesis is the conformational conversion from an α-helix or random coil structure to a β-sheet structure During the second growth phase, soluble oligomers and protofibrils with molecular weights of approximately 20 kDa (1–2 nm) (low MW) for tetramers and 56–60 kDa (3–5 nm) (high MW) for dodecamers are formed, respectively [38]. Finally, in the stationary phase, amyloid fibrils are formed [39]. Additionally, fibrils can form aggregates through secondary mechanisms, which can be classified based on their dependence or independence on monomer concentration, such as elongation and secondary nucleation (dependent) or fragmentation (independent) (Figure 1) [40,41,42]. In secondary nucleation, the surfaces of fibrils catalyze the formation of new low-molecular-weight (LMW) oligomers. Once a critical fibril concentration is reached, this mechanism becomes the primary source of oligomers, surpassing the primary nucleation process [40]. This positive feedback leads to the rapid, exponential accumulation of toxic oligomers [43,44,45]. Recent findings have shown that the severity of the disease correlates with the levels of soluble Aβ oligomers [46,47], which are recognized as the main neurotoxic agents. These oligomers disrupt various cellular processes through multiple mechanisms [39,48,49]. They can bind to numerous membrane receptors, disrupting normal signaling pathways, and also form pore-like structures in the neuronal membrane, altering its permeability and integrity (as reviewed in [39]). As a result, the dysregulation of Ca2+ homeostasis, mitochondrial dysfunction, reactive oxygen species (ROS) generation, reduced ATP levels, and tau phosphorylation occurs, ultimately leading to synaptic dysfunction and neuronal loss. Consequently, the amyloid cascade hypothesis has been revised to the “oligomer hypothesis” [50].
Based on the current understanding of amyloidogenesis, researchers are focusing on several strategies for treating AD. These strategies include the following: (i) Reducing Aβ production through α-secretase activators, and β-secretase (BACE1) and γ-secretase inhibitors. (ii) Enhancing Aβ clearance via active or passive immunization, receptor-mediated clearance, modulation of APP-tau interaction, and targeting Aβ and neuroinflammation. A significant breakthrough was achieved through the passive immunization approach, which involves the direct administration of antibodies. This method has proven to be highly effective, with the antibodies being well tolerated and considered safe due to their high specificity [51,52]. Two monoclonal antibodies, Aducanumab and Lecanemab, received approval from the US Food and Drug Administration (FDA) on 7 July 2021, and 22 January 2023, respectively, for the treatment of AD [53,54]. (iii) Inhibiting Aβ oligomerization with amyloid inhibitors, which can be categorized into various types: small organic molecules (both natural and synthetic), carbon-based nanomaterials (such as fullerenes, carbon nanotubes, graphene oxide, and carbon dots), nanoparticles (including polymer, metallic, and lipid-based types), and metal chelators (such as iron, copper, and zinc ion chelators). For more information on AD treatment strategies, see references [15,51,52,53,54,55,56,57]. An underexplored approach in studying and understanding the molecular mechanisms of oligomer formation, the kinetics of aggregation, and designing potential modulators of fibrillogenesis involves promoters, rather than inhibitors, of Aβ aggregation [58,59,60,61,62,63,64].

3. Structural Characteristics and Forms of Monomers, Dimers, and Oligomers

The size of soluble oligomers ranges from 10 to 100 kDa and they exhibit significant heterogeneity and dynamic behavior. According to the widely recognized protein folding funnel model, the conformational space of a folding protein or peptide is vast [65,66,67,68,69,70]. Unfolded soluble monomers occupy the broad top of the funnel, possessing the highest energy and the ability to adopt numerous conformations. Folded intermediates or partially folded states, as well as oligomers and native forms of the monomer, are lower in energy and correspond to local minima within the energy landscape. Amorphous aggregates are located at one of the narrower bottoms of the funnel, characterized by deep energy minima, while the global free energy minimum, at the narrowest part of the funnel, is occupied by amyloid fibrils [66]. Fibrilization begins with an ensemble of unfolded monomer conformations and rapidly proceeds along various pathways toward the global minimum, where distinct amyloid polymorphs occupy closely positioned local minima. The absolute free energy minimum is associated with amyloid crystals [68,69]. The exact mechanisms and conformational transitions from disordered monomers through oligomers to fibrils remain challenging to understand, due to the heterogeneity, metastability, and dynamic nature of oligomers. Additionally, fibrillogenesis is strongly influenced by factors such as temperature, concentration, and the homogeneity of the starting monomer structures. Various Aβ oligomer structures, including U-shaped, S-shaped, LS-shaped, and those with two- or three-fold topologies, have been identified through solution and solid-state NMR and cryo-electron microscopy, with some originating from human sources, as shown in Figure 2 [71,72,73,74,75,76].
At the structural level, the predominant secondary structure of amyloid fibrils is the cross β-sheet, where the Aβ backbone is oriented orthogonally to the fibril axis [77,78,79,80,81]. In the oligomer mixture of the on-pathway, β-sheet structure is commonly observed. In solution, the Aβ monomer typically adopts an unfolded conformation rather than any distinct secondary structure (Figure 2). The primary structure of Aβ1–42 is divided into five regions: the N-terminus, also known as the hydrophilic or metal-binding region (D1 to Q15); the central hydrophobic core (CHC), spanning residues K16 to A21; the loop or central hydrophilic region (E22 to K28); the second hydrophobic region (G29 to M35); and the C-terminal region (V36 to A42) (Figure 2) [56].

4. Interactions Leading to and Stabilizing Aβ Aggregation

It has been found that within the β-hairpin structured monomers that form Aβ oligomers, intrastrand hydrogen bonds occur between the β-strand regions, specifically between I31 and V36 [82]. During the transformation from β-hairpin monomers to β-sheet secondary structure, these intrastrand bonds must be broken, and new interstrand hydrogen bonds are formed between adjacent peptide sequences [83]. This shift from intra- to interstrand hydrogen bonds is essential for the oligomerization process, resulting in the formation of the cross β-sheet structure. It has also been found that during the Aβ association process, monomers primarily interact with polar surfaces such as mica, while hydrophobic surfaces such as graphite disrupt the oligomer structure and act as templates for fibrillization [84]. Furthermore, it is thought that the initial interaction responsible for the β-sheet structure in mature fibrils involves a hydrophobic contact between F19 from the central hydrophobic core (CHC) of one peptide and L34 from the second hydrophobic region of another monomer [85,86,87]. A critical interaction that stabilizes the turn in the cross β-structure is the salt bridge between D23 and K28 [88,89,90]. During oligomerization, it is believed that monomer oligomerization occurs via parallel stacking in the axial direction of fibril extension. However, published fibril structures suggest that the dimeric Aβ1–42 unit, composed of two S-shaped monomers arranged in a C2 symmetric “ying-yang” manner, is involved in fibril growth [74,75,76]. It has been identified that the key interactions stabilizing all types of quaternary fibril structures involve the side chains of M35 and one or more residues, such as I31, I32, and M35 from one monomer, interacting with G37, G39, and V29 from a second monomer [91,92,93,94,95]. In the U-shaped Aβ17–42 form, K28-D23 salt bridges serve as key stabilizing interactions [96]. In contrast, the S-shaped Aβ11–42 and LS-shaped Aβ1–42 fibrils are stabilized by salt bridges between the positively charged NH3+ group of the K28 side chain and the negatively charged COO group of A42 [73,75]. Additionally, in the LS-shaped form, the N-terminal and C-terminal regions are reinforced by E11-H6/H13 hydrogen bonds, which play a crucial role in fibril stability. Recent studies emphasize the importance of the hydrophobic core formed by F4, L34, and V36, along with the K28-A42 salt bridge, in stabilizing the LS-shaped Aβ fibril [75].

5. Molecular Dynamics Simulations: Challenges and Prospects

Understanding the mechanisms by which small organic molecules inhibit oligomerization and designing new inhibitors is a complex challenge for traditional experimental techniques. This difficulty arises from the vast conformational space occupied by Aβ monomers and oligomers, the metastable nature of oligomers during fibrillization, and their dynamic structural changes. Computational approaches, such as molecular dynamics simulations, serve as valuable complementary tools in this research field.
Molecular dynamics (MD) is a powerful structure-based drug design (SBDD) method that enhances our understanding of the properties, functions, and mechanisms of action or inhibition of biomacromolecules and other systems of interest. MD is currently the only applicable method capable of simulating the internal motions of biomolecules at the atomic level, providing insights that are otherwise unattainable using conventional experimental techniques [97,98]. Integrating experimental data with MD simulations leads to a more comprehensive understanding of molecular processes and mechanisms. In conventional or classical MD (cMD), also known as all-atom MD, each atom is represented as a particle that has position and velocity, evolving according to Newton’s laws of motion. The interactions between atoms are calculated using force fields (FFs), which are mathematical models that determine the potential energy of a system by accounting for bonded and non-bonded interactions. These parameters can be derived from experimental data or computed using quantum mechanics. Since different molecular systems, such as proteins, nucleotides, and small organic molecules, require specialized force fields, no single universal FF applies to all cases. Well-established force fields for proteins include Amber ff14S [99], GROMOS versions 54A7 and 54B7 [100], CHARMM [101], and OPLS3 [102]. Most of these force fields are optimized for structured proteins. However, since Aβ peptides belong to the category of intrinsically disordered proteins (IDPs), specialized FFs have been developed to describe their dynamic behavior more accurately. A recent review by Boopathi and colleagues extensively discusses various force fields tailored for IDPs [56]. Some of these include ff99IDPs, ff14IDPSFF [103], ff14IDPs [104], ff03 CMAP [105], ff14SB [99], CHARMM36m [106], and CHARMM36IDPSFF [107], all designed for better modeling of misfolded and folded protein structures [108,109,110,111,112,113].
While specialized force fields have been developed for IDPs, their suitability depends on system design and simulation conditions, such as sampling methods and water models. Furthermore, comparisons between different FFs are based on diverse criteria, as shown in Table 1. Various software packages also recommend specific FFs and water models for IDPs. For instance, the latest version of Amber24 suggests using the ff99SD force field with the general-purpose four-point OPC water model [114,115].
Carballo-Pacheco and Strodel highlighted that the error in calculating NMR observables is so significant that determining which force field best represents experimental data remains challenging [116]. Additionally, multiple studies present conflicting results regarding the most accurate force field for modeling IDPs and protein aggregation. However, newer force fields generally outperform older ones, reflecting continuous advancements in their accuracy over time.
Table 1. Comparison of force fields for IDPs based on various criteria and system specifications. Force fields are marked in bold if the authors conclude that their findings align with experimental data.
Table 1. Comparison of force fields for IDPs based on various criteria and system specifications. Force fields are marked in bold if the authors conclude that their findings align with experimental data.
Examined FFsCriteria for
Comparison
Studied
System/s
Sampling Method/
Water Model
Ref.
AMBER94, AMBER96, AMBER99, AMBER99-ILDN, AMBER03, AMBER12SB, AMBER14SB, GROMOS43a1, GROMOS43a2GROMOS45a3, GROMOS53a5, GROMOS53a6, GROMOS54a7, CHARM22*, CHARRMM36, CHARRMM36m, and OPLS-AAglobal reaction coordinates, secondary structure content, fibril population and formation100 dimeric Aβ16–22cMD/TIP3P for all FFs, except for GROMOS FFs―SPC, CHARM22*―mTIP3P, and OPLS―TIP4P[117]
ff99SB, ff14SB, FF14SB_IDPs, CHARMM36, CHARMM36mchemical shifts, secondary structure content, nonbonded energy component, Enonbondedmonomeric Aβ42cMD and REMD/TIP3P[118]
ff99IDPs, ff14IDPs, ff14IDPSFF, ff03w, CHARMM36m, and CHARMM22*chemical shits, J-couplings, global reaction coordinates, secondary structure contentRS-peptide, HEWL19, HIV-rev, Aβ40, Aβ42, phosphodiesterase-γ, CspTm, and ubiquitincMD/for ff99IDPs, ff14IDPs, ff14IDPSFF―TIP3P, for ff03w―TIP4P-2005; and for CHARMM FFs―CHARMM-modified TIP3P [113]
Gromos54a7, OPLS-AA, AMBER03ws, CHARMM22*, and AMBER99SB*ILDNoligomer formation kinetics, in terms of dissociation constant, KD and ∆G, and collision acceptance probabilitymonomeric forms and six monomers of each of the three peptides: Aβ16–22, one non-amyloidogenic mutant (F19V, F20V), and the Aβ16–22 (F19L) mutant, which exhibits rapid fibrillogenesis SPC for Gromos54a7, TIP4P/2005 for AMBER03ws, TIP4P for OPLS-AA, TIP4P-Ew for CHARMM22* and AMBER99SB*ILDN [119]
OPLS, AMBER99SB, AMBER99SB*ILDN, AMBER99SBILDN-NMR and CHARMM22*local NMR observables, including chemical shifts, J-couplings, and residual dipolar couplings (RDCs)monomeric Aβ1–42REMD/TIP4P-Ew for AMBER99SB, AMBER99SB*ILDN, AM-BER99SBILDN-NMR and CHARMM22*, TIP3P for OPLS[116]
Abbreviation: REMD―replica exchange molecular dynamics.
The derivatives of the potential energy functions correspond to the forces in Newton’s equations of motion, which are solved iteratively for each atom in time steps of femtoseconds. While cMD enables highly precise calculations, its simulations are limited in timescale, typically reaching only microseconds to milliseconds, even with modern computational advancements. However, given the stochastic nature of biological processes, this duration is often insufficient to achieve comprehensive Boltzmann sampling or to capture critical changes in biologically relevant phenomena such as protein folding and amyloid formation. To overcome these drawbacks, including timescale limitations and sampling deficiencies, various advanced sampling methods have been developed using different techniques [120,121,122,123,124,125,126]. A comprehensive review and classification of the numerous sampling methods was published by Zuckerman [124]. Here, we will briefly describe some of them, considering the scope and purpose of this review. In this review, we will briefly highlight some of these methods, focusing on the scope and objectives of our discussion. For in-depth discussions on cMD and advanced MD techniques, refer to several articles [121,124,125,126,127,128,129,130,131,132].
The generalized-ensemble algorithm enables a system to bypass energy barriers of local minima and sample a much broader conformational space [133,134,135]. This is achieved by replacing the Boltzmann probability weight factor, which decreases exponentially with the energy barrier height, with non-Boltzmann factors, ensuring a uniform distribution of potential energy space. As a result, a random walk in temperature leads to a random walk in potential energy. One example of this approach is simulated tempering (ST), which requires the initial determination of probability weight factors, a computationally demanding task [136,137]. This issue is bypassed in the widely used replica exchange method (REM) [138,139,140,141,142] (also known as parallel tempering [133,140,142,143], replica Monte Carlo [90,91,138], and multiple Markov chain method [139]), where the probability weight factor is predefined. The method involves simulating multiple non-interacting replicas independently, each representing a copy of the original system in the canonical ensemble at different fixed temperatures. Periodically, replicas at neighboring temperatures are swapped based on the Metropolis criterion. The replica exchange method applied to MD, called REMD (replica exchange molecular dynamics), was developed by Sugita and Okamoto [144]. In REMD, in addition to considering coordinates and potential energy functions, momenta are accounted for, and the velocities of all atoms in a pair of replicas are uniformly rescaled. This method is particularly suitable for parallel computing, as each temperature exchange must be synchronized. Therefore, the application of this method to complex biological systems with explicit solvent molecules is constrained by computational power, as these systems have an immense number of degrees of freedom, necessitating the simulation of a large number of replicas simultaneously on a vast, homogeneous computer cluster. To address this challenge, various variants and optimizations of the replica exchange method have been developed [145,146,147,148,149,150,151,152,153,154,155,156].
A relatively recent method based on the generalized-ensemble algorithm is simulated tempering distributed replica (STDR), which is designed for use on distributed or heterogeneous platforms [130,157]. The method requires minimal initial simulation, unlike ST, and is not affected by replica synchronization. Additionally, it does not require a fixed number of temperatures, unlike RE. However, when using generalized-ensemble methods, it is important to consider the potential loss of dynamic information typical of cMD, due to non-physical transitions and changes that may occur during simulations.
One approach to overcome the limitations of cMD is to reduce the degrees of freedom. It is clear that increasing the system size raises the degrees of freedom, which, in turn, limits the duration of simulations. Coarse-grained (CG) modeling, where a single particle represents a group of atoms instead of individual atoms, significantly reduces the degrees of freedom, allowing simulation timescales to be extended by 2–4-times compared to atomistic representations, though this comes at the cost of reduced accuracy. Various coarse-grained models (CGMs) exist, ranging from one-bead to united-atom representations [158,159,160]. For example, a coarse-grained model for proteins is implemented in the MARTINI force field [161]. Another method for reducing degrees of freedom is by representing the solvent implicitly using implicit solvent-based approaches [162].
An advanced strategy for improving conformational sampling and overcoming the timescale limitations of cMD involves modifying the potential energy in the force field. Techniques like conformational flooding involve incorporating a flooding potential into the force field, which destabilizes the local energy minima, facilitating the escape from initial energy wells without altering the reaction pathway [163,164]. Another method is metadynamics, also known as the local elevation method. This technique involves adding a Gaussian-type biasing potential to the Hamiltonian, pushing the system away from previously visited conformational states [165,166,167,168,169]. It requires prior selection of significant coordinates to which repulsive terms are applied. Metadynamics is particularly effective for systems with a smooth energy landscape. Another technique is umbrella sampling, where an artificial biasing potential, referred to as the umbrella, is incorporated into the potential energy function. This bias directs the sampling towards less favorable conformational regions. The umbrella potential acts as a compensating function, typically concentrating the sampling within a specific region, known as a window. By shifting the focus through multiple simulations, the entire conformational space can be explored. Prior knowledge of the reaction coordinate is essential for constructing the umbrella potential, as the method heavily relies on the chosen potential and the number of windows [170,171]. In accelerated MD (aMD), the potential energy landscape of the original system is modified by introducing a bias potential, which raises the energy minima of the wells and reduces the energy barriers between different states. The advantage of this approach is that it does not require prior knowledge of the system’s potential energy landscape, eliminating the need for defining the reaction coordinate in advance. Initially introduced by Voter under the name “hyperdynamics,” the goal was to shorten the residence time of the system within the potential minima and overcome energy barriers, thereby enabling the study of long-term system behavior [172,173]. The early version of this bias potential was time-consuming and limited to small systems. Later, Hamelberg and colleagues developed a bias potential that could be applied to larger systems, such as biomolecules [174]. The key innovation of this bias potential is that it preserves the overall shape of the system’s potential energy surface. To ensure that thermodynamic quantities, like free energy changes derived from the modified potential, match those from the unmodified potential, a reweighting step is necessary. However, in systems with large biases, typical of large systems, the exponential reweighting factor causes the ensemble averages to rely on only a few configurations with high weights, reducing the data points and affecting the accuracy of thermodynamic quantities [175]. To address this issue, variants of aMD have been developed, including replica exchange accelerated molecular dynamics (REXAMD) [176], selective aMD [177], and others [178,179,180].
Despite the development of numerous advanced techniques, and while they are often highly effective, it is important to remember that there is no single universal method that can serve as an ideal solution for the timescale limitations of cMD in all cases [181,182].
Alongside the development of MD sampling methods, force fields and algorithms have been continuously improved (some examples can be found [183,184,185]). Additionally, computational power continues to expand. These advancements have contributed to extending the timescale of simulations, making MD simulations more accessible and indispensable for uncovering the functions and mechanisms of biological systems. For example, the latest version of the supercomputer designed specifically for conventional MD simulations of biological systems, Anton 3, can perform cMD simulations of systems with a million atoms at over 100 microseconds per day [186]. The third-fastest supercomputer on the Top-500 June 2023 list [187], LUMI (Rmax (PFlop/s) 309.10), can simulate 80 nanoseconds per day for a system with over a million atoms [188]. The second-fastest supercomputer on the same list is capable of simulating 8.30 ns per day for a system of 1.6 billion atoms, allowing for cellular-scale simulations [184]. In addition to the evolution of supercomputers, the distributed computing project Folding@home (FAH or F@h) was developed [189,190]. In 2020, during the pandemic, over a million citizen scientists contributed their computing resources to simulate the SARS-CoV-2 proteome, making the FAH platform one of the largest computational resources in the world. It was the first time it crossed the exascale computing barrier and successfully simulated 0.1 s of the viral proteome [191].
In 2012, Lemkul and Bevan reviewed the pioneering in silico studies on amyloid beta inhibitors, including molecular docking and molecular dynamics simulations of both Aβ fragments and the full-length peptide [192]. They also highlighted the limitations of existing methods used for Aβ systems and fibrils, such as the restricted timescales of production dynamics, the use of implicit solvent models, and the constraints of scoring functions.
Over the past decade, only a few articles have reviewed the applications of SBDD and methods, including molecular docking and/or molecular dynamics simulations, for various Aβ inhibitors [56,193,194,195]. To our knowledge, none have focused on the mechanistic insights gained from MD simulations of endogenous compounds and repurposed drugs, which is the main objective of this review.
Endogenous compounds—naturally synthesized within the human body—present a promising approach to disease treatment. This strategy is considered one of the most natural, leading to minimal or no toxicity and being well tolerated with few or no side effects. These compounds have demonstrated superior therapeutic outcomes. Another effective approach is drug repurposing, discovering new therapeutic applications for existing drugs.
The following sections focus on MD studies that uncover the mechanisms of these two classes of small molecules. Given the limited research over the past decade specifically targeting the toxic Aβ1–42 form, we also include MD studies on the Aβ1–40 isoform of the peptide.

6. Endogenous Compounds Inhibiting Aβ

6.1. Dopamine (DA) and Norepinephrine (NE)

Dopamine (DA) and norepinephrine (NE) (also known as noradrenaline) (Figure 3) belong to the catecholamine family and function as neurotransmitters and neuromodulators. In the brain, DA is involved in motor control, executive functions, and the motivational aspects of reward-driven behavior [196]. NE mobilizes the brain and body for action, playing a vital role in maintaining alertness, attention, mood, memory, and stress response [197]. Additionally, it is crucial for supporting contextual and spatial memory [198].
A study of the anti-amyloidogenic and fibril-destabilizing effects of anti-Parkinsonian agents found that DA dose-dependently inhibits the formation and extension of Aβ1–40 and Aβ1–42 fibrils, as well destabilizing preformed fibrils [199]. Another study revealed that catechol derivatives, including catechol, DA, L-DOPA, NE, epinephrine, and L-tyrosine, primarily inhibit nucleation rather than the elongation of Aβ fibrilization in the presence and absence of a model biomembrane [200]. Among the studied compounds, DA exhibited the strongest inhibitory activity, followed by L-DOPA and NE, in preventing Aβ1–40 and Aβ1–42 fibril formation.
A REMD study on Aβ1–40 protofibril (Table 2) revealed that DA preferably binds to β-2 BS, a β-sheet of IIGLMVG (residues 31–37) from the second hydrophobic region of Aβ, as well as to the N-terminal site, which presents disordered tails [201]. The study found that the ligand significantly impacted the oligomer’s double-layered structure.
In a comprehensive study conducted by Chen et al., the disruptive mechanisms of DA were investigated using cMD simulations of protonated DA+ molecules at different molecular ratios with preformed pentameric Aβ protofibrils, as well as REMD simulations of DA+ with Aβ dimers (Table 2) [202]. The authors found that in the systems with low molar ratios (1:1 and 2:1), DA+ molecules destabilized Aβ protofibrils in a dose-dependent manner by inserting into the F4-L34-V36 core region and disrupting intra- and interchain K28-A42 salt bridges. In the 1:1 system, DA+ molecules primarily bound to the turn-1 region (H6-H13), while those in the 2:1 system also interacted with the F4-L34-V36 core region, the N-terminal (D1-R5), the turn-2 region (F20-D23), and the C-terminal (I41 and A42). These interactions occurred through hydrogen bonding with residues D1, E3, H6, D7, E11, H13, Q15, E22, D23 and A42; salt bridge formation with D7 and E11 (for the 1:1 system) and E11, E22 and D23 (for the 2:1 system); cation-π interactions with R5; and π-π stacking with F4, H6, H13, H14, F19 and F20.
At higher concentrations (molar ratio 10:1), the protofibril structure remained intact as the DA+ molecules predominantly bound to the outer surface of Aβ protofibril, limiting its flexibility (Table 2). However, when 10 DA+ molecules were replaced with deprotonated DA0—which exists at low concentrations under physiological pH—DA0 molecules preferentially bound to the inner surface of the protofibril, particularly within the F4-L34-V36 hydrophobic core. The deprotonated DA0 molecules interacted with protonated DA+ molecules via π-π stacking, enhancing their binding to the inner surface and ultimately exerting a disruptive effect on the protofibril structure. Additionally, the authors examined the inhibitory effects of DA+ molecules on Aβ dimerization (Table 2) [202].
The mechanism of Aβ dimer aggregation inhibition and fibril destabilization by NE was studied using REMD and cMD simulations, respectively (Table 2) [203]. During dimer fibrillization, NE was found to decrease β-sheet content while increasing α-helix, coil, and turn content. Five main binding sites (BSs) were identified: (1) 3EFRHD7, (2) 10YEVHH14, (3) 16KLVFFA21, (4) 31IIGLMV36, and (5) 39VVIA42. Among these, the third and fourth sites were the most favorable, as they had the lowest binding energy. NE molecules primarily interacted with hydrophobic residues I41, I31, and L17, as well as aromatic residues Y10, F4, and F20, through hydrophobic and stacking interactions, respectively. Additionally, hydrogen bonds and cation-π interactions contributed to NE binding to Aβ peptides. Hydrogen bonds were mainly formed with negatively charged aspartic and glutamic acid residues (D1, E3, D7, E11, E22, and D23), while cation-π interactions occurred with the positively charged side chain of R5.
Table 2. Summary of the technical data and key conclusions from the reviewed MD simulations. Most studies applied conventional MD, while some utilized alternative methods, which are specified accordingly.
Table 2. Summary of the technical data and key conclusions from the reviewed MD simulations. Most studies applied conventional MD, while some utilized alternative methods, which are specified accordingly.
FF/Water ModelDuration per System, nsAβ Length/PDB ID/Type
(Monomer/Dimer/
(Proto-)Fibril)
Inhibitor * Main FindingsRef.
REMD/GROMOS 57a7/SPC50 per replica1–40/2LMN/decamer, protofibrilDApreferably binds to β-2 and N-terminal;
significantly affects the oligomer’s double-layer structure
[201]
AMBER99SBILDN/TIP3P5 × 500 per system1–42/5OQV/pentamer(1) 5 DA+* (molecular ratio 1:1)
(2) 10 DA+ (2:1)
(3) 50 DA+ (10:1)
(4) 40 DA+ + 10 DA0 (10:1)
(1) binds to H6-E11 (turn-1 region); weak disruptive effect;
(2) binds to H6-E11, the F4-L34-V36 hydrophobic core, the turn-2 region (F20, E22 and D23), and the C-terminal residues I41 and A42; stronger disruptive effect
(3) binds mainly to the outer surface, decreasing the flexibility of the Aβ protofibril and stabilizing it
(4) DA0 molecules bind to the inner surface of the protofibril, primarily to the F4-L34-V36 hydrophobic core; π-π stacking with DA+ increases their inner surface binding, resulting in a disruptive effect of the DA0 and DA+ mixture
[202]
REMD/AMBER99SBILDN/TIP3P950 per replica (48 replicas)(1) Aβ1–42/1IYT/two monomers placed in three orientations: parallel, antiparallel and perpendicular10 DA+ (molar ratio 5:1)inhibits the dimerization[202]
REMD/AMBER99SBILDN/TIP3P300 per replica (54 replicas)1–42/1Z0Q/dimer20 NE (molar ratio 10:1)suppresses and reduces the interpeptide β-sheet content;
five dominant BS were identified;
main contact residues: hydrophobic interactions with L17, I31, and I41; π-π stacking with Y4, F10, and F20; H-bonds with D1, E3, D7, E11, E22, and D23; cation-π interactions with R5
[203]
AMBER99SBILDN/TIP3P2 × 1000
(2 × 1 µs)
1–42/5OQV/pentamer100 NEβ-sheet content decreased, while coil content increased; the number of fibril H-bonds decreased; H-bonds formed with D1, A2, D23, and A42; destabilizes the preformed fibril[203]
AMBER99SBILDN/TIP3P3 × 500 (per system)1–42/5OQV/tetramer(1) 20 NE+ (molar ration 5:1)
(2) 20 NE0 (molar ration 5:1)
(1) destabilizes the protofibril; primarily disrupts the D1-H14 region; decreases β-sheet content; increases the kink angle around Y10; disrupts H6-E11 H-bonds and K28-A42 salt bridges; forms H-bonds with E3, D7, E11, Q15, E22, and D23; engages in π-π stacking with H6, H13, and F20;
(2) disrupts the protofibril; decreases β-sheet content in dispersed regions, including A2–H6, A21–G22, S26, and M35–V36; primarily forms hydrophobic interactions and π-π stacking with F4, R5, H6, Y10, H13, Q15, F20 and L34.
[204]
AMBER99SBILDN/TIP3P3 × 5001–42/5OQV/tetramer40 SER (molar ratio 10:1)disrupts fibrils by decreasing β-sheet content at the N-terminal (D1-Y10); main contact residues include F4, H6, Y10, H13, Q15, and L34; dominant interactions involve π-π stacking with F4, H6, Y10, and H13; disrupts fibril-stabilizing contacts A2-V36 and F4-L34 at core-1; as well as salt bridges between L34-A42.[205]
AMBER99SBILDN/TIP3P3 × 5001–42/5OQV/tetramer40 MEL (molar ratio 10:1)disrupts fibrils by decreasing β-sheet content at the N-terminal (D1–Y10) and C-terminal (D23–A42); two BSs were identified: (1) at the N-terminal (F4, H6, Y10, H13, H14, Q15, L17, and F19) and (2) at the C-terminal (N27, I31, I32, L34, and V36); dominant interactions include π-π stacking with F4, H6, Y10, H13, H14, and F19, as well as hydrophobic interactions with N27, I31, I32, L34 and V36; disrupts fibril-stabilizing contacts A2-V36 and F4-L34 at core-1, L17-I31 at core-2, and I32-M35 at core-3, along with salt bridges between L34 andA42[205]
AMBER14SB/TIP3P(1) 500
(2) 500
(3) 3 × 500
16–22/random pentamer(1) 20 ATP (molar ratio 4:1)
(2) 25 ATP (molar ratio 5:1)
(3) 30 ATP (molar ratio 6:1)
with increasing ATP concentration, (i) β-sheet content decreased, while turn, bend, and coil contents increased, preventing oligomerization; (ii) the ATP-F π-π stacking disrupted the F-F interpeptide interactions; (iii) peptide-peptide H-bonds decreased, while ATP-peptide H-bonds increased;[206]
AMBER99SB-ILDN/TIP3P
AMBER-FB15/TIP3P-FB
2 × 500 (for each FF)16–22/random pentamer30 ATP (molar ratio 6:1)β-sheet content decreases sharply; interpeptide F-F interactions are reduced; peptide-peptide H-bonds decrease abruptly, preventing β-sheets formation;[206]
AMBER14SB/TIP3P
AMBER99SB-ILDN/TIP3P
AMBER-FB15/TIP3P-FB
500 (for each FF)16–22/dimer16 ATP (molar ratio 8:1)unfavored dimerization[206]
AMBER14SB/TIP3P
AMBER99SB-ILDN/TIP3P
AMBER-FB15/TIP3P-FB
50016–22/prefibrillar pentamer150 ATP (molecular ration 30:1)destabilizes β-sheet content and promotes disaggregation; completely disaggregates fibrils in the AMBER14SB and AMBER-FB15 FFs; a significant decreases inter-peptide H-bonds[206]
DA—dopamine; NE—norepinephrine; SER—serotonin; MEL—melatonin; ATP—adenosine triphosphate; BS—binding site; * number of molecules is specified when different from 1; +—protonated form; 0—unprotonated form.
In the fibril system, NE molecules led to a decrease in β-sheet content while increasing coil content [203]. NE molecules remodeled the Aβ fibril structure through hydrogen bonding with D1, A2, D23, and A42 residues, ultimately destabilizing the preformed fibril.
cMD simulations of the protonated (NE+) and deprotonated (NE0) forms of norepinephrine revealed that both forms disrupt the LS-shaped tetrameric protofibril through distinct mechanisms (Table 2) [204]. The protonated form, which is the predominant form at physiological pH, destabilized the entire protofibril by primarily disrupting the N-terminal region (D1–H14), reducing its β-sheet content, and increasing the kink angle around Y10. It formed hydrogen bonds with E3, D7, E11, Q15, E22, and D23, as well as π-π stacking interactions with H6, H13, and F20. As a result, the H-bonds between H6 and E11 along with the salt bridges between K28 and A42—both critical for fibril stabilization—were disrupted. The deprotonated form disrupted the protofibril by reducing β-sheet content in several scattered regions, including A2–H6, A21–E22, S26, and M35–V36. Its interactions were primarily hydrophobic and involved π-π stacking with F4, H6, Y10, H13, Q15, F20, and L34. The interactions with F4 and L34 likely interfered with the essential stabilizing contacts between these residues, ultimately disrupting the F4–L34–V36 hydrophobic core.

6.2. Serotonin (SER) and Melatonin (MEL)

Serotonin (SER) (Figure 3), an indoleamine neurotransmitter involved in regulating mood, cognition, learning, and memory, and its derivative melatonin (MEL), a hormone essential for controlling the circadian rhythm or sleep–wake cycle, are both synthesized in the brain from the amino acid tryptophan. In patients with Alzheimer’s disease (AD), cortical serotonin levels are reduced due to acute tryptophan depletion, contributing to cognitive decline [207,208]. Additionally, SER exhibits neuroprotective effects against Aβ toxicity and effectively inhibits Aβ fibrillization [209,210]. MEL levels decline with age, with a more pronounced reduction observed in AD patients [211,212,213,214]. MEL has been shown to exert a neuroprotective effect by inhibiting Aβ aggregation; it directly interacts with Aβ peptides, prevents amyloid beta formation, and promotes Aβ fibril dissociation [215,216,217,218].
A recent study by Gong et al. investigated the molecular-level effects of SER and MEL on LS-shaped Aβ fibrils (Table 2) [205]. Interestingly, both indoleamine derivatives exhibited distinct mechanisms of Aβ fibril destabilization. SER primarily binds to the N-terminal region (D1-Y10), disrupting its β-sheet content mainly through π-π stacking interactions with F4, H6, Y10, and H13. As a result, it interferes with the fibril-stabilizing interactions between A2 and V36, as well as between F4 and L34. The binding site (BS) residues of SER include F4, H6, Y10, H13, Q15, and L34, as SER mainly interacts with inner-surface-oriented residues whose side chains are embedded within the Aβ protofibril, such as F4, H6, H13, Q15, and L34. Furthermore, SER insertion induces distortion of the C-terminal region (mainly D23-A42) and disrupts intra- and interchain K28–A42 salt bridges, with a stronger effect on intrachain interactions.
In contrast, MEL binds to two binding sites on the L-shaped Aβ protofibril—one at the N-terminal segment, comprising residues F4, H6, Y10, H13, H14, Q15, L17, and F19, and the other at the C-terminal region, including residues N27, I31, I32, L34, and V36. MEL disrupts the β-sheet structure in both the N-terminal (D1-Y10) and C-terminal (D23-A42) segments, exhibiting a stronger destabilizing effect on the entire protofibril. It interferes with fibril-stabilizing interactions at three hydrophobic cores: A2-V36 and F4-L34 at core-1, L17-I31 at core-2, and I32-M35 at core-3. The impact of MEL on L34-A42 inter- and intrachain salt bridges is more significant than that of SER, with a greater effect on intrachain interactions. MEL primarily engages in two types of intermolecular interactions: π-π stacking with aromatic amino acids in the N-terminal region, such as F4, H6, Y10, H13, H14, and F19, and hydrophobic contacts with C-terminal residues, including N27, I31, I32, L34, and V36.

6.3. Adenosine Triphosphate (ATP)

Adenosine triphosphate (ATP) (Figure 3) is recognized as the primary energy source for various processes in living cells. Experimental studies have shown that ATP acts as a biological hydrotrope—an amphiphilic molecule that, unlike conventional surfactants, exhibits low aggregation cooperativity and functions at molar concentrations. ATP prevents the formation of peptide aggregates and dissolves pre-existing ones at the millimolar scale [219]. Recently, an experimental study revealed that ATP induced amorphous aggregation of Aβ by increasing its dynamics [220].
Coskuner and Murray studied the effect of ATP on Aβ fibrils, both computationally and in vitro, finding that ATP reduced Aβ misfolding at physiological intracellular concentrations of approximately 500 µM [221]. The in silico calculations, which included initial molecular docking followed by short 10 ns cMD simulations on the Aβ9–40 fibril and thermodynamic analyses, showed that ATP primarily interacted with Y10 and S26, with Y10 playing a crucial role in ATP binding to the fibril.
A thorough investigation by Pal and Paul examined the inhibitory effect of ATP on the central hydrophobic core (CHC) Aβ16–22 of the Aβ peptide using three different force fields: AMBER14SB, AMBER99SB-ILDN, and AMBER-FB15. The peptide was considered as a dimer, random pentamer, and prefibrillar pentamer at different molecular ratios (Table 2) [206]. Trajectory analyses of the random pentamer simulations across the three FFs showed similar results in the presence of ATP. Specifically, ATP inhibited the oligomerization of Aβ peptides at millimolar concentrations, as indicated by a decrease in β-sheet content, peptide–peptide hydrogen bonds, and intrapeptide F-F interactions. Meanwhile, ATP-F π-π stacking interactions and ATP-peptide hydrogen bonds increased. The MD simulations of the dimer systems across all three FFs showed that ATP inhibited dimerization. Additionally, ATP destabilized the β-sheet content and induced disaggregation in pre-formed fibril systems simulated with AMBER14SB, AMBER99SB-ILDN, and AMBER-FB15 FFs. Notably, in the AMBER14SB and AMBER-FB15 FFs, the fibrils were completely disaggregated. These results were supported by a significant decrease in interpeptide H-bonds.

7. Repurposed Drugs Inhibiting Aβ

7.1. Propafenone (PPF)

To discover novel anti-amyloidogenic agents, one approach involves repurposing existing approved drugs, while another strategy focuses on screening compounds with high similarity to known inhibitors. Combining these methods, Ngo et al. conducted a similarity search of FDA-approved drugs with over 80% similarity to curcumin, followed by molecular docking against the Aβ9–40 fibril, a dodecamer with two-fold symmetry [222]. They then performed cMD simulations (Table 3) of the selected drugs, followed by absolute binding energy calculations, which identified Propafenone (PPF) (Figure 4), an anti-arrhythmic drug, as the most promising Aβ inhibitor. Further trajectory analyses revealed that PPF was located near the turn in the lower layer of the dodecameric fibril, primarily interacting with the hydrophobic residues. Additionally, it was found that β-content decreased in the presence of PPF, potentially leading to fibril degradation. Experimental validation further confirmed the anti-amyloidogenic activity of PPF against Aβ40 (IC50(PPF) = 1.8 µM, IC50 (curcumin) = 2.7 µM) and Aβ42 (IC50(PPF) = 3.9 µM, IC50 (curcumin) = 6.9 µM), as well as its antioxidant properties and ability to reduce Aβ40-induced cytotoxicity. PPF also exhibited protective effects against Aβ40- and Aβ42-induced toxicity in SH-SY5Y cells.

7.2. Carbenoxolone (CBX)

Carbenoxolone (3-O-(3-Carboxypropanoyl)-3β-hydroxy-11-oxoolean-12-en-30-oic acid) (Figure 4) is a natural derivative of an 18-glycyrrhetinic acid with a steroid-like structure found in liquorice root. It is a drug with pleotropic pharmacologic activity [223] and has been clinically used to treat various types of ulceration and inflammation. Known as a moderately potent gap junction blocker [224], the compound is water-soluble and can cross the blood–brain barrier (BBB). Carbenoxolone has shown nootropic and neuroprotective effects [225,226]. It reversibly inhibits the activity of 11β-hydroxysteroid dehydrogenase type 1 (11β-HSD1), which catalyzes the conversion of inactive cortisone to the active glucocorticoid cortisol in the brain, resulting in memory improvement [227]. Sharma et al. studied the inhibitory effect of CBX on Aβ aggregation, both experimentally and computationally [228]. The authors initially predicted binding sites on the monomeric and fibrillar forms of Aβ1–42 using the CASTp and RaptorX servers, followed by molecular docking calculations. In the final step of their computational analysis, they ran classical MD simulations of the best-docked complexes between CBX and both forms of Aβ1–42 forms for 10 ns (Table 3). They found that CBX reduced the secondary structure of both Aβ1–42 forms, decreasing α-helix and β-sheet content in the monomeric form and β-sheet content in the fibrillar form, while increasing unstructured content. The ligand formed contacts with F4, R5, H6, Y10, V12, H14, Q15, K16 V18, F19, and A30 in the monomer and with L17, V18, F19, F20, A21, D23, K38, L34, V36, V40, and I32 in the fibril chains. The compound also formed hydrogen bonds with R5, Q15, and F4 in the monomer and with F19 and D23 in the fibril, disrupting the salt bridge between D23 and K38.

7.3. Doxycycline (DXC)

In vivo and in vitro studies of the anticancer agent Iododoxorubicin have indicated its anti-amyloidogenic properties [229,230,231,232]. However, due to its significant toxicity, particularly cardiotoxicity, efforts have focused on identifying structurally similar compounds that retain anti-fibrillogenic activity while exhibiting a safer toxicological profile and have already been tested in clinical settings. Tetracyclines, which contain a polycyclic conjugated moiety similar to that of iododoxorubicin, have been shown to exhibit anti-amyloidogenic properties against different proteins, including Aβ [233,234,235,236,237,238,239,240,241,242,243]. Doxycycline (DXC) has been found to inhibit Aβ42 aggregation and disassemble mature amyloid fibrils [235,239,240].
Table 3. Repurposed drugs. A summary of the technical data and key findings from the reviewed MD simulations. Conventional MD was used in most studies, with alternative methods indicated in some cases.
Table 3. Repurposed drugs. A summary of the technical data and key findings from the reviewed MD simulations. Conventional MD was used in most studies, with alternative methods indicated in some cases.
FF/Water ModelDuration per System, nsAβ Length/PDB ID/Type
(Monomer/Dimer/
(Proto-)Fibril)
InhibitorMain FindingsRef.
GROMOS96 43a1/SPC4 × 259–40/2LMN/dodecamer
1–40/2M4J/nonamer
PPFMainly forms contacts with hydrophobic residues; β-content decreases; [222]
GROMOS96/SPC10(1) Aβ1–42/1IYT/monomer
(2)Aβ17–42/2BEG/pentamer, protofibril
CBX(1) reduces α-helix and β-sheet secondary structure; increases unstructured contend; forms contacts with F4, R5, H6, Y10, V12, H14, Q15, K16 V18, F19, and A30; forms H-bonds with R5, Q15, and F4;
(2) reduces β-sheet secondary structure; increases unstructured content; forms contacts with L17, V18, F19, F20, A21, D23, K38, L34, V36, V40, and I32; forms H-bonds with F19 and D23; disrupts the salt bridge between D23 and K38.
[228]
aMD/ff14SB/TIP3P3 × 1000 (1 µs)(1) Aβ11–42/2MXU/pentamer
(2) Aβ1–42/5OQV/pentamer
5 DXC (molar ratio 1:1)(1) destabilizes the hydrophobic core (N15-A30); three main BSs: (i) near the M35 side chain; (ii) between I32 and L34; and (iii) between L17 and F19;
(2) two BSs were identified: (i) near E1, V39, and I41; and (ii) at K16, V18, and F20.
[244]
The anti-amyloidogenic mechanism of DXC against two Aβ fibril polymorph forms, S-shaped and LS-shaped, was studied using accelerated MD (Table 3) [244]. It was found that the S-shaped pentameric fibril was partially destabilized, affecting the hydrophobic core consisting of residues N15–A30. Three main binding sites (BSs) were identified. The first was located near the side chain of M35, oriented toward the outer face of the fibril, where DXC interacted with the fibril backbone via hydrogen bonds and with M35 side chains through hydrophobic contacts. The second binding site was located between I32 and L34 on the solvent-exposed face of the larger hydrophobic core (N15–L34), where DXC formed hydrophobic contacts with the I32 and L34 side chains, as well as hydrogen bonds with the fibril backbone. The third binding site was identified between residues L17 and F19 at the solvent-accessible region of the hydrophobic core (N15–L34), where DXC was stabilized through hydrophobic interactions with L17 and F19. For the second fibril form, these sites are hidden; therefore, two different BSs were identified. The first was near the N-terminal of the LS-shaped fibril, where DXC engaged in polar interactions with E1 and hydrophobic contacts with V39 and I41. The second binding site consisted of residues K16, V18, and F20.

8. Mechanisms of Aβ Inhibition in Drug Discovery

Identifying Aβ inhibitors through structure-based drug design (SBDD) using molecular docking and molecular dynamics simulations is highly challenging. Typically, virtual screening through molecular docking, pharmacophore searches, or similarity searches can uncover potential novel Aβ inhibitors. Subsequent MD simulations can further strengthen these predictions, while in vitro and in vivo assays can validate them. Recently, a rapid in vivo pipeline for screening small-molecule inhibitors of Aβ aggregation was introduced [245]. Targeting fibrillogenesis is difficult due to its inherent complexity; the wide range of monomeric, dimeric, and oligomeric conformations with multiple binding sites; and the potential for false-positive predictions from computational approximations. However, a promising example was reviewed here [222]. Effective inhibitors must disrupt key fibril-stabilizing interactions or those essential for aggregation, as discussed in Section 3. Usually, the main affected secondary structure is the β-sheet content, an outstanding mark of Aβ oligomeric structures.
Moreover, both SBDD techniques and ligand-based drug design (LBDD) methods (a recent comprehensive review on the LBDD QSAR approach was published by us [246] and can be applied to known Aβ inhibitors) can be effectively employed in drug design for hit-to-lead optimization, based on a validated protocol, model, or pipeline of known Aβ inhibitors. These predictions can then be experimentally validated.
Mechanisms of Aβ inhibition can be categorized into the three main types: (1) Early-stage inhibition—This targets primary nucleation, i.e., the process of dimerization, where a small ligand hinders the interaction between two monomers. This interference can lead to an irregular or disordered nucleus or even prevent nucleation entirely (see Figure 1). (2) Middle-stage inhibition—This targets the formation of oligomers by interfering with the growth of pre-existing soluble low-molecular-weight oligomers (LMWO), such as dimers, trimers, and tetramers, which act as nucleation sites (see Figure 1). This mechanism can also affect secondary nucleation. (3) Disruption of existing oligomers—This strategy destabilizes already-formed LMWO or high-molecular-weight oligomers (HMWO), leading to the formation of irregular nuclei and non-fibrillar LMWO (see Figure 1). As a result, the aggregation process is diverted off-pathway, potentially yielding irregular β-aggregates or even complete disaggregation.
Unveiling the precise mechanisms of Aβ inhibition can be achieved through well-planned and properly conducted MD simulations, as outlined in the next section. Furthermore, these findings can be validated through in vitro and/or in vivo experiments and may also help guide future studies to confirm the proposed mechanisms.

9. Challenges and Perspectives in MD Simulations

The complexity of fibrillization leads to fibrils with diverse morphologies, further complicating the process of uncovering and gaining deep insights into the inhibition mechanisms of small molecules. Additionally, selecting force fields that best mimic the fibrillization process, along with an appropriate water model, is crucial. A third challenge lies in the availability of computational resources. To achieve a comprehensive and optimal understanding of the inhibition mechanisms of a given anti-amyloidogenic inhibitor using conventional MD simulations, the following key considerations should be taken into account: (1) The full length of the peptide should be considered, regardless of whether the simulation is performed on a monomeric peptide or a later-stage fibrillization unit such as a dimer, tetramer, or pentamer. (2) Simulations on protofibrils should be conducted for different morphological forms, such as U-, S-, and SL-shaped fibrils, to gain a thorough understanding of all currently known fibril structures. (3) Careful and appropriate selection of force fields for MD simulations is essential, and utilizing multiple force fields is preferable and recommended. (4) The water model should be carefully chosen to align with the selected force field for consistency and accuracy. (5) The molar ratio of the inhibitor to the peptide should be consistent with experimental data and, if possible, should mimic brain concentrations. Multiple molar ratios should also be considered. (6) In addition to studying the inhibitor of interest, a positive and negative control inhibitor, as well as a simulation without any inhibitor, should be included using the same MD simulation protocol. Additionally, the protonation state of the molecules should be considered at physiological pH. (7) The length of production dynamics should be sufficiently long to capture relevant structural changes, depending on computational resources and simulation time. (8) To gain a deeper understanding of the inhibition mechanism of a given anti-amyloidogenic compound, a comprehensive analysis should be performed, considering key stabilizing interactions of specific (proto-)fibril forms as well as the critical protein secondary structure element for fibril formation—the cross β-sheet content. (9) Finally, multiple simulation runs with different initial velocities should be performed to generate more reliable and statistically representative trajectories.

Funding

This research was funded by the Centre of Excellence for Informatics and Information and Communication Technologies, supported by the Science and Education for Smart Growth Operation Program, Bulgarian Ministry of Education and Science Grant No. BG05M2OP001-1.001-0003.

Data Availability Statement

Not applicable.

Acknowledgments

During the preparation of this work, the author used an AI tool for English editing. After using this tool, the author reviewed and edited the content as needed and took full responsibility for the content of the publication.

Conflicts of Interest

The author declares no conflict of interest.

Abbreviations and Acronyms

aaamino acids
ADAlzheimer’s disease
AICD APP intracellular domain
aMDaccelerated molecular dynamics
APPamyloid-β (A4) precursor protein
ATPadenosine triphosphate
BACE1β—secretase 1, β- site amyloid precursor protein (APP) cleavaging enzyme 1
BBBblood–brain barrier
BSbinding site
CBXCarbenoxolone
CGcoarse-grained
CGMcoarse-grained model
cMDclassical/conventional molecular dynamics
CTFC-terminal fragment of APP
DAdopamine
FADfamilial Alzheimer’s disease
FFforce field
IDPintrinsically disordered proteins
LMWlow molecular weight
LMWOlow molecular weight oligomers
MDmolecular dynamics
MELmelatonin
NEnorepinephrine
NMRnuclear magnetic resonance
PPFPropafenone
REreplica exchange
REMreplica exchange method
REMDreplica exchange molecular dynamics
REXAMDreplica exchange accelerated molecular dynamics
ROSreactive oxygen species
SADsporadic Alzheimer’s disease
SERserotonin
SPCsimple point charge water model
STDRsimulated tempering distributed replica
TIP3Ptransferable intermolecular potential 3-point water model
TMDtransmembrane domain

References

  1. GBD 2019 Collaborators Global Mortality from Dementia: Application of a New Method and Results from the Global Burden of Disease Study 2019. A&D Transl. Res. Clin. Interv. 2021, 7, e12200. [CrossRef]
  2. World Health Organization. Dementia. 2023. Available online: https://www.who.int/news-room/fact-sheets/detail/dementia (accessed on 20 March 2023).
  3. Ghiso, J.; Frangione, B. Amyloidosis and Alzheimer’s Disease. Adv. Drug Deliv. Rev. 2002, 54, 1539–1551. [Google Scholar] [CrossRef]
  4. Hamley, I.W. The Amyloid Beta Peptide: A Chemist’s Perspective. Role in Alzheimer’s and Fibrillization. Chem. Rev. 2012, 112, 5147–5192. [Google Scholar] [CrossRef] [PubMed]
  5. Akiyama, H. Inflammation and Alzheimer’s Disease. Neurobiol. Aging 2000, 21, 383–421. [Google Scholar] [CrossRef] [PubMed]
  6. Cvetković-Dožić, D.; Skender-Gazibara, M.; Dožić, S. Neuropathological Hallmarks of Alzheimer’s Disease. Arch. Oncol. 2001, 9, 195–199. [Google Scholar]
  7. Zhang, W.; Pei, J.; Lai, L. Computational Multitarget Drug Design. J. Chem. Inf. Model. 2017, 57, 403–412. [Google Scholar] [CrossRef] [PubMed]
  8. Polanco, J.C.; Li, C.; Bodea, L.-G.; Martinez-Marmol, R.; Meunier, F.A.; Götz, J. Amyloid-β and Tau Complexity—Towards Improved Biomarkers and Targeted Therapies. Nat. Rev. Neurol. 2018, 14, 22–39. [Google Scholar] [CrossRef] [PubMed]
  9. Bartus, R.T. On Neurodegenerative Diseases, Models, and Treatment Strategies: Lessons Learned and Lessons Forgotten a Generation Following the Cholinergic Hypothesis. Exp. Neurol. 2000, 163, 495–529. [Google Scholar] [CrossRef]
  10. Terry, A.V.; Buccafusco, J.J. The Cholinergic Hypothesis of Age and Alzheimer’s Disease-Related Cognitive Deficits: Recent Challenges and Their Implications for Novel Drug Development. J. Pharmacol. Exp. Ther. 2003, 306, 821–827. [Google Scholar] [CrossRef]
  11. Winek, K.; Soreq, H.; Meisel, A. Regulators of Cholinergic Signaling in Disorders of the Central Nervous System. J. Neurochem. 2021, 158, 1425–1438. [Google Scholar] [CrossRef] [PubMed]
  12. Selkoe, D.J.; Hardy, J. The Amyloid Hypothesis of Alzheimer’s Disease at 25 Years. EMBO Mol. Med. 2016, 8, 595–608. [Google Scholar] [CrossRef]
  13. Du, X.; Wang, X.; Geng, M. Alzheimer’s Disease Hypothesis and Related Therapies. Transl. Neurodegener. 2018, 7, 2. [Google Scholar] [CrossRef]
  14. Naseri, N.N.; Wang, H.; Guo, J.; Sharma, M.; Luo, W. The Complexity of Tau in Alzheimer’s Disease. Neurosci. Lett. 2019, 705, 183–194. [Google Scholar] [CrossRef] [PubMed]
  15. Uddin, M.S.; Kabir, M.T.; Rahman, M.S.; Behl, T.; Jeandet, P.; Ashraf, G.M.; Najda, A.; Bin-Jumah, M.N.; El-Seedi, H.R.; Abdel-Daim, M.M. Revisiting the Amyloid Cascade Hypothesis: From Anti-Aβ Therapeutics to Auspicious New Ways for Alzheimer’s Disease. Int. J. Mol. Sci. 2020, 21, 5858. [Google Scholar] [CrossRef] [PubMed]
  16. Arnsten, A.F.T.; Datta, D.; Del Tredici, K.; Braak, H. Hypothesis: Tau Pathology Is an Initiating Factor in Sporadic Alzheimer’s Disease. Alzheimer’s & Dement. 2021, 17, 115–124. [Google Scholar] [CrossRef]
  17. Hardy, J.A.; Higgins, G.A. Alzheimer’s Disease: The Amyloid Cascade Hypothesis. Science 1992, 256, 184–185. [Google Scholar] [CrossRef] [PubMed]
  18. Leroy, K.; Ando, K.; Laporte, V.; Dedecker, R.; Suain, V.; Authelet, M.; Héraud, C.; Pierrot, N.; Yilmaz, Z.; Octave, J.-N.; et al. Lack of Tau Proteins Rescues Neuronal Cell Death and Decreases Amyloidogenic Processing of APP in APP/PS1 Mice. Am. J. Pathol. 2012, 181, 1928–1940. [Google Scholar] [CrossRef] [PubMed]
  19. Borchelt, D.R.; Thinakaran, G.; Eckman, C.B.; Lee, M.K.; Davenport, F.; Ratovitsky, T.; Prada, C.-M.; Kim, G.; Seekins, S.; Yager, D.; et al. Familial Alzheimer’s Disease–Linked Presenilin 1 Variants Elevate Aβ1–42/1–40 Ratio In Vitro and In Vivo. Neuron 1996, 17, 1005–1013. [Google Scholar] [CrossRef] [PubMed]
  20. Kabir, M.T.; Uddin, M.S.; Setu, J.R.; Ashraf, G.M.; Bin-Jumah, M.N.; Abdel-Daim, M.M. Exploring the Role of PSEN Mutations in the Pathogenesis of Alzheimer’s Disease. Neurotox. Res. 2020, 38, 833–849. [Google Scholar] [CrossRef]
  21. Strittmatter, W.J.; Saunders, A.M.; Schmechel, D.; Pericak-Vance, M.; Enghild, J.; Salvesen, G.S.; Roses, A.D. Apolipoprotein E: High-Avidity Binding to Beta-Amyloid and Increased Frequency of Type 4 Allele in Late-Onset Familial Alzheimer Disease. Proc. Natl. Acad. Sci. USA 1993, 90, 1977–1981. [Google Scholar] [CrossRef]
  22. Saunders, A.M.; Strittmatter, W.J.; Schmechel, D.; St. George-Hyslop, P.H.; Pericak-Vance, M.A.; Joo, S.H.; Rosi, B.L.; Gusella, J.F.; Crapper-MacLachlan, D.R.; Alberts, M.J.; et al. Association of Apolipoprotein E Allele Ε4 with Late-onset Familial and Sporadic Alzheimer’s Disease. Neurology 1993, 43, 1467. [Google Scholar] [CrossRef]
  23. Bertram, L.; Tanzi, R.E. Thirty Years of Alzheimer’s Disease Genetics: The Implications of Systematic Meta-Analyses. Nat. Rev. Neurosci. 2008, 9, 768–778. [Google Scholar] [CrossRef]
  24. Mawuenyega, K.G.; Sigurdson, W.; Ovod, V.; Munsell, L.; Kasten, T.; Morris, J.C.; Yarasheski, K.E.; Bateman, R.J. Decreased Clearance of CNS β-Amyloid in Alzheimer’s Disease. Science 2010, 330, 1774. [Google Scholar] [CrossRef]
  25. Yang, L.-B.; Lindholm, K.; Yan, R.; Citron, M.; Xia, W.; Yang, X.-L.; Beach, T.; Sue, L.; Wong, P.; Price, D.; et al. Elevated β-Secretase Expression and Enzymatic Activity Detected in Sporadic Alzheimer Disease. Nat. Med. 2003, 9, 3–4. [Google Scholar] [CrossRef]
  26. Dawkins, E.; Small, D.H. Insights into the Physiological Function of the β-amyloid Precursor Protein: Beyond Alzheimer’s Disease. J. Neurochem. 2014, 129, 756–769. [Google Scholar] [CrossRef]
  27. Halim, A.; Brinkmalm, G.; Rüetschi, U.; Westman-Brinkmalm, A.; Portelius, E.; Zetterberg, H.; Blennow, K.; Larson, G.; Nilsson, J. Site-Specific Characterization of Threonine, Serine, and Tyrosine Glycosylations of Amyloid Precursor Protein/Amyloid β-Peptides in Human Cerebrospinal Fluid. Proc. Natl. Acad. Sci. USA 2011, 108, 11848–11853. [Google Scholar] [CrossRef] [PubMed]
  28. Zhang, Y.; Thompson, R.; Zhang, H.; Xu, H. APP Processing in Alzheimer’s Disease. Mol. Brain 2011, 4, 3. [Google Scholar] [CrossRef]
  29. Zerr, I. (Ed.) Understanding Alzheimer’s Disease; InTech: London, UK, 2013; ISBN 978-953-51-1009-5. [Google Scholar]
  30. Chen, G.; Xu, T.; Yan, Y.; Zhou, Y.; Jiang, Y.; Melcher, K.; Xu, H.E. Amyloid Beta: Structure, Biology and Structure-Based Therapeutic Development. Acta Pharmacol. Sin. 2017, 38, 1205–1235. [Google Scholar] [CrossRef]
  31. Gandy, S. The Role of Cerebral Amyloid β Accumulation in Common Forms of Alzheimer Disease. J. Clin. Investig. 2005, 115, 1121–1129. [Google Scholar] [CrossRef] [PubMed]
  32. Wilkins, H.M.; Swerdlow, R.H. Amyloid Precursor Protein Processing and Bioenergetics. Brain Res. Bull. 2017, 133, 71–79. [Google Scholar] [CrossRef] [PubMed]
  33. Urban, A.S.; Pavlov, K.V.; Kamynina, A.V.; Okhrimenko, I.S.; Arseniev, A.S.; Bocharov, E.V. Structural Studies Providing Insights into Production and Conformational Behavior of Amyloid-β Peptide Associated with Alzheimer’s Disease Development. Molecules 2021, 26, 2897. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, Y.; McLaughlin, R.; Goodyer, C.; LeBlanc, A. Selective Cytotoxicity of Intracellular Amyloid β Peptide1–42 through P53 and Bax in Cultured Primary Human Neurons. J. Cell Biol. 2002, 156, 519–529. [Google Scholar] [CrossRef] [PubMed]
  35. Owen, M.C.; Kulig, W.; Poojari, C.; Rog, T.; Strodel, B. Physiologically-Relevant Levels of Sphingomyelin, but Not GM1, Induces a β-Sheet-Rich Structure in the Amyloid-β(1-42) Monomer. Biochim. Biophys. Acta (BBA)-Biomembr. 2018, 1860, 1709–1720. [Google Scholar] [CrossRef]
  36. Gargari, S.A.; Barzegar, A. Simulations on the Dual Effects of Flavonoids as Suppressors of Aβ42 Fibrillogenesis and Destabilizers of Mature Fibrils. Sci. Rep. 2020, 10, 16636. [Google Scholar] [CrossRef]
  37. Du, W.-J.; Guo, J.-J.; Gao, M.-T.; Hu, S.-Q.; Dong, X.-Y.; Han, Y.-F.; Liu, F.-F.; Jiang, S.; Sun, Y. Brazilin Inhibits Amyloid β-Protein Fibrillogenesis, Remodels Amyloid Fibrils and Reduces Amyloid Cytotoxicity. Sci. Rep. 2015, 5, 7992. [Google Scholar] [CrossRef] [PubMed]
  38. Fu, Z.; Aucoin, D.; Ahmed, M.; Ziliox, M.; Van Nostrand, W.E.; Smith, S.O. Capping of Aβ42 Oligomers by Small Molecule Inhibitors. Biochemistry 2014, 53, 7893–7903. [Google Scholar] [CrossRef] [PubMed]
  39. Huang, Y.; Liu, R. The Toxicity and Polymorphism of β-Amyloid Oligomers. Int. J. Mol. Sci. 2020, 21, 4477. [Google Scholar] [CrossRef] [PubMed]
  40. Cohen, S.I.A.; Linse, S.; Luheshi, L.M.; Hellstrand, E.; White, D.A.; Rajah, L.; Otzen, D.E.; Vendruscolo, M.; Dobson, C.M.; Knowles, T.P.J. Proliferation of Amyloid-β42 Aggregates Occurs through a Secondary Nucleation Mechanism. Proc. Natl. Acad. Sci. USA 2013, 110, 9758–9763. [Google Scholar] [CrossRef] [PubMed]
  41. Linse, S. Monomer-Dependent Secondary Nucleation in Amyloid Formation. Biophys. Rev. 2017, 9, 329–338. [Google Scholar] [CrossRef] [PubMed]
  42. Scheidt, T.; Łapińska, U.; Kumita, J.R.; Whiten, D.R.; Klenerman, D.; Wilson, M.R.; Cohen, S.I.A.; Linse, S.; Vendruscolo, M.; Dobson, C.M.; et al. Secondary Nucleation and Elongation Occur at Different Sites on Alzheimer’s Amyloid-β Aggregates. Sci. Adv. 2019, 5, eaau3112. [Google Scholar] [CrossRef]
  43. Ferrone, F.A.; Hofrichter, J.; Eaton, W.A. Kinetics of Sickle Hemoglobin Polymerization. J. Mol. Biol. 1985, 183, 611–631. [Google Scholar] [CrossRef]
  44. Knowles, T.P.J.; Waudby, C.A.; Devlin, G.L.; Cohen, S.I.A.; Aguzzi, A.; Vendruscolo, M.; Terentjev, E.M.; Welland, M.E.; Dobson, C.M. An Analytical Solution to the Kinetics of Breakable Filament Assembly. Science 2009, 326, 1533–1537. [Google Scholar] [CrossRef]
  45. Cohen, S.I.A.; Vendruscolo, M.; Dobson, C.M.; Knowles, T.P.J. Nucleated Polymerization with Secondary Pathways. II. Determination of Self-Consistent Solutions to Growth Processes Described by Non-Linear Master Equations. J. Chem. Phys. 2011, 135, 065106. [Google Scholar] [CrossRef] [PubMed]
  46. Tabaton, M.; Nunzi, M.G.; Xue, R.; Usiak, M.; Autiliogambetti, L.; Gambetti, P. Soluble Amyloid β-Protein Is a Marker of Alzheimer Amyloid in Brain But Not in Cerebrospinal Fluid. Biochem. Biophys. Res. Commun. 1994, 200, 1598–1603. [Google Scholar] [CrossRef] [PubMed]
  47. Frackowiak, J.; Zoltowska, A.; Wisniewski, H.M. Non-Fibrillar β-Amyloid Protein Is Associated with Smooth Muscle Cells of Vessel Walls in Alzheimer Disease. J. Neuropathol. Exp. Neurol. 1994, 53, 637–645. [Google Scholar] [CrossRef] [PubMed]
  48. Shankar, G.M.; Li, S.; Mehta, T.H.; Garcia-Munoz, A.; Shepardson, N.E.; Smith, I.; Brett, F.M.; Farrell, M.A.; Rowan, M.J.; Lemere, C.A.; et al. Amyloid-β Protein Dimers Isolated Directly from Alzheimer’s Brains Impair Synaptic Plasticity and Memory. Nat. Med. 2008, 14, 837–842. [Google Scholar] [CrossRef] [PubMed]
  49. Choi, Y.J.; Chae, S.; Kim, J.H.; Barald, K.F.; Park, J.Y.; Lee, S.-H. Neurotoxic Amyloid Beta Oligomeric Assemblies Recreated in Microfluidic Platform with Interstitial Level of Slow Flow. Sci. Rep. 2013, 3, 1921. [Google Scholar] [CrossRef] [PubMed]
  50. Hardy, J.; Selkoe, D.J. The Amyloid Hypothesis of Alzheimer’s Disease: Progress and Problems on the Road to Therapeutics. Science 2002, 297, 353–356. [Google Scholar] [CrossRef]
  51. Decourt, B.; Boumelhem, F.; Pope, E.D.; Shi, J.; Mari, Z.; Sabbagh, M.N. Critical Appraisal of Amyloid Lowering Agents in AD. Curr. Neurol. Neurosci. Rep. 2021, 21, 39. [Google Scholar] [CrossRef] [PubMed]
  52. Panza, F.; Lozupone, M.; Logroscino, G.; Imbimbo, B.P. A Critical Appraisal of Amyloid-β-Targeting Therapies for Alzheimer Disease. Nat. Rev. Neurol. 2019, 15, 73–88. [Google Scholar] [CrossRef] [PubMed]
  53. Arndt, J.W.; Qian, F.; Smith, B.A.; Quan, C.; Kilambi, K.P.; Bush, M.W.; Walz, T.; Pepinsky, R.B.; Bussière, T.; Hamann, S.; et al. Structural and Kinetic Basis for the Selectivity of Aducanumab for Aggregated Forms of Amyloid-β. Sci. Rep. 2018, 8, 6412. [Google Scholar] [CrossRef]
  54. Wu, W.; Ji, Y.; Wang, Z.; Wu, X.; Li, J.; Gu, F.; Chen, Z.; Wang, Z. The FDA-Approved Anti-Amyloid-β Monoclonal Antibodies for the Treatment of Alzheimer’s Disease: A Systematic Review and Meta-Analysis of Randomized Controlled Trials. Eur. J. Med. Res. 2023, 28, 544. [Google Scholar] [CrossRef]
  55. Pagano, K.; Tomaselli, S.; Molinari, H.; Ragona, L. Natural Compounds as Inhibitors of Aβ Peptide Aggregation: Chemical Requirements and Molecular Mechanisms. Front. Neurosci. 2020, 14, 619667. [Google Scholar] [CrossRef]
  56. Boopathi, S.; Poma, A.B.; Garduño-Juárez, R. An Overview of Several Inhibitors for Alzheimer’s Disease: Characterization and Failure. Int. J. Mol. Sci. 2021, 22, 10798. [Google Scholar] [CrossRef] [PubMed]
  57. Li, J.; Liao, W.; Huang, D.; Ou, M.; Chen, T.; Wang, X.; Zhao, R.; Zhang, L.; Mei, L.; Liu, J.; et al. Current Strategies of Detecting Aβ Species and Inhibiting Aβ Aggregation: Status and Prospects. Coord. Chem. Rev. 2023, 495, 215375. [Google Scholar] [CrossRef]
  58. Zhao, Y.; Rao, P.P.N. Small Molecules N-Phenylbenzofuran-2-Carboxamide and N -Phenylbenzo[b]Thiophene-2-Carboxamide Promote Beta-Amyloid (Aβ42) Aggregation and Mitigate Neurotoxicity. ACS Chem. Neurosci. 2023, 14, 4185–4198. [Google Scholar] [CrossRef] [PubMed]
  59. Bieschke, J.; Herbst, M.; Wiglenda, T.; Friedrich, R.P.; Boeddrich, A.; Schiele, F.; Kleckers, D.; Lopez Del Amo, J.M.; Grüning, B.A.; Wang, Q.; et al. Small-Molecule Conversion of Toxic Oligomers to Nontoxic β-Sheet–Rich Amyloid Fibrils. Nat. Chem. Biol. 2012, 8, 93–101. [Google Scholar] [CrossRef]
  60. Luo, J.; Yu, C.-H.; Yu, H.; Borstnar, R.; Kamerlin, S.C.L.; Gräslund, A.; Abrahams, J.P.; Wärmländer, S.K.T.S. Cellular Polyamines Promote Amyloid-Beta (Aβ) Peptide Fibrillation and Modulate the Aggregation Pathways. ACS Chem. Neurosci. 2013, 4, 454–462. [Google Scholar] [CrossRef] [PubMed]
  61. De Almeida, N.E.C.; Do, T.D.; Tro, M.; LaPointe, N.E.; Feinstein, S.C.; Shea, J.-E.; Bowers, M.T. Opposing Effects of Cucurbit[7]Uril and 1,2,3,4,6-Penta-O-Galloyl-β-d-Glucopyranose on Amyloid β25–35 Assembly. ACS Chem. Neurosci. 2016, 7, 218–226. [Google Scholar] [CrossRef] [PubMed]
  62. Sonzini, S.; Stanyon, H.F.; Scherman, O.A. Decreasing Amyloid Toxicity through an Increased Rate of Aggregation. Phys. Chem. Chem. Phys. 2017, 19, 1458–1465. [Google Scholar] [CrossRef] [PubMed]
  63. Liu, H.; Qian, C.; Yang, T.; Wang, Y.; Luo, J.; Zhang, C.; Wang, X.; Wang, X.; Guo, Z. Small Molecule-Mediated Co-Assembly of Amyloid-β Oligomers Reduces Neurotoxicity through Promoting Non-Fibrillar Aggregation. Chem. Sci. 2020, 11, 7158–7169. [Google Scholar] [CrossRef]
  64. Mohamed, T.; Gujral, S.S.; Rao, P.P.N. Tau Derived Hexapeptide AcPHF6 Promotes Beta-Amyloid (Aβ) Fibrillogenesis. ACS Chem. Neurosci. 2018, 9, 773–782. [Google Scholar] [CrossRef] [PubMed]
  65. Bryngelson, J.D.; Onuchic, J.N.; Socci, N.D.; Wolynes, P.G. Funnels, Pathways, and the Energy Landscape of Protein Folding: A Synthesis. Proteins 1995, 21, 167–195. [Google Scholar] [CrossRef]
  66. Onuchic, J.N.; Wolynes, P.G. Theory of Protein Folding. Curr. Opin. Struct. Biol. 2004, 14, 70–75. [Google Scholar] [CrossRef] [PubMed]
  67. Dill, K.A.; Chan, H.S. From Levinthal to Pathways to Funnels. Nat. Struct. Mol. Biol. 1997, 4, 10–19. [Google Scholar] [CrossRef]
  68. Reynolds, N.P.; Adamcik, J.; Berryman, J.T.; Handschin, S.; Zanjani, A.A.H.; Li, W.; Liu, K.; Zhang, A.; Mezzenga, R. Competition between Crystal and Fibril Formation in Molecular Mutations of Amyloidogenic Peptides. Nat. Commun. 2017, 8, 1338. [Google Scholar] [CrossRef]
  69. Adamcik, J.; Mezzenga, R. Amyloid Polymorphism in the Protein Folding and Aggregation Energy Landscape. Angew. Chem. Int. Ed. 2018, 57, 8370–8382. [Google Scholar] [CrossRef] [PubMed]
  70. Ke, P.C.; Zhou, R.; Serpell, L.C.; Riek, R.; Knowles, T.P.J.; Lashuel, H.A.; Gazit, E.; Hamley, I.W.; Davis, T.P.; Fändrich, M.; et al. Half a Century of Amyloids: Past, Present and Future. Chem. Soc. Rev. 2020, 49, 5473–5509. [Google Scholar] [CrossRef] [PubMed]
  71. Lührs, T.; Ritter, C.; Adrian, M.; Riek-Loher, D.; Bohrmann, B.; Döbeli, H.; Schubert, D.; Riek, R. 3D Structure of Alzheimer’s Amyloid-β(1–42) Fibrils. Proc. Natl. Acad. Sci. USA 2005, 102, 17342–17347. [Google Scholar] [CrossRef] [PubMed]
  72. Schmidt, M.; Rohou, A.; Lasker, K.; Yadav, J.K.; Schiene-Fischer, C.; Fändrich, M.; Grigorieff, N. Peptide Dimer Structure in an Aβ(1–42) Fibril Visualized with Cryo-EM. Proc. Natl. Acad. Sci. USA 2015, 112, 11858–11863. [Google Scholar] [CrossRef]
  73. Xiao, Y.; Ma, B.; McElheny, D.; Parthasarathy, S.; Long, F.; Hoshi, M.; Nussinov, R.; Ishii, Y. Aβ(1–42) Fibril Structure Illuminates Self-Recognition and Replication of Amyloid in Alzheimer’s Disease. Nat. Struct. Mol. Biol. 2015, 22, 499–505. [Google Scholar] [CrossRef]
  74. Colvin, M.T.; Silvers, R.; Ni, Q.Z.; Can, T.V.; Sergeyev, I.; Rosay, M.; Donovan, K.J.; Michael, B.; Wall, J.; Linse, S.; et al. Atomic Resolution Structure of Monomorphic Aβ42 Amyloid Fibrils. J. Am. Chem. Soc. 2016, 138, 9663–9674. [Google Scholar] [CrossRef] [PubMed]
  75. Gremer, L.; Schölzel, D.; Schenk, C.; Reinartz, E.; Labahn, J.; Ravelli, R.B.G.; Tusche, M.; Lopez-Iglesias, C.; Hoyer, W.; Heise, H.; et al. Fibril Structure of Amyloid-β(1–42) by Cryo–Electron Microscopy. Science 2017, 358, 116–119. [Google Scholar] [CrossRef]
  76. Wälti, M.A.; Ravotti, F.; Arai, H.; Glabe, C.G.; Wall, J.S.; Böckmann, A.; Güntert, P.; Meier, B.H.; Riek, R. Atomic-Resolution Structure of a Disease-Relevant Aβ(1–42) Amyloid Fibril. Proc. Natl. Acad. Sci. USA 2016, 113. [Google Scholar] [CrossRef] [PubMed]
  77. Eanes, E.D.; Glenner, G.G. X-RAY DIFFRACTION STUDIES ON AMYLOID FILAMENTS. J. Histochem. Cytochem. 1968, 16, 673–677. [Google Scholar] [CrossRef] [PubMed]
  78. Kirschner, D.A.; Abraham, C.; Selkoe, D.J. X-Ray Diffraction from Intraneuronal Paired Helical Filaments and Extraneuronal Amyloid Fibers in Alzheimer Disease Indicates Cross-Beta Conformation. Proc. Natl. Acad. Sci. USA 1986, 83, 503–507. [Google Scholar] [CrossRef]
  79. Sunde, M.; Blake, C. The Structure of Amyloid Fibrils by Electron Microscopy and X-Ray Diffraction. In Advances in Protein Chemistry; Elsevier: Amsterdam, The Netherlands, 1997; Volume 50, pp. 123–159. ISBN 978-0-12-034250-1. [Google Scholar]
  80. Serpell, L.C. Alzheimer’s Amyloid Fibrils: Structure and Assembly. Biochim. Biophys. Acta (BBA)-Mol. Basis Dis. 2000, 1502, 16–30. [Google Scholar] [CrossRef]
  81. Makin, O.S.; Serpell, L.C. Structures for Amyloid Fibrils. FEBS J. 2005, 272, 5950–5961. [Google Scholar] [CrossRef] [PubMed]
  82. Ahmed, M.; Davis, J.; Aucoin, D.; Sato, T.; Ahuja, S.; Aimoto, S.; Elliott, J.I.; Van Nostrand, W.E.; Smith, S.O. Structural Conversion of Neurotoxic Amyloid-β1–42 Oligomers to Fibrils. Nat. Struct. Mol. Biol. 2010, 17, 561–567. [Google Scholar] [CrossRef] [PubMed]
  83. Härd, T. Protein Engineering to Stabilize Soluble Amyloid β-protein Aggregates for Structural and Functional Studies. FEBS J. 2011, 278, 3884–3892. [Google Scholar] [CrossRef]
  84. Kowalewski, T.; Holtzman, D.M. In Situ Atomic Force Microscopy Study of Alzheimer’s β-Amyloid Peptide on Different Substrates: New Insights into Mechanism of β-Sheet Formation. Proc. Natl. Acad. Sci. USA 1999, 96, 3688–3693. [Google Scholar] [CrossRef]
  85. Adler, J.; Scheidt, H.A.; Krüger, M.; Thomas, L.; Huster, D. Local Interactions Influence the Fibrillation Kinetics, Structure and Dynamics of Aβ(1–40) but Leave the General Fibril Structure Unchanged. Phys. Chem. Chem. Phys. 2014, 16, 7461–7471. [Google Scholar] [CrossRef]
  86. Korn, A.; McLennan, S.; Adler, J.; Krueger, M.; Surendran, D.; Maiti, S.; Huster, D. Amyloid β (1–40) Toxicity Depends on the Molecular Contact between Phenylalanine 19 and Leucine 34. ACS Chem. Neurosci. 2018, 9, 790–799. [Google Scholar] [CrossRef] [PubMed]
  87. Bonhommeau, S.; Talaga, D.; Hunel, J.; Cullin, C.; Lecomte, S. Tip-Enhanced Raman Spectroscopy to Distinguish Toxic Oligomers from Aβ 1–42 Fibrils at the Nanometer Scale. Angew. Chem. Int. Ed. 2017, 56, 1771–1774. [Google Scholar] [CrossRef] [PubMed]
  88. Qiang, W.; Yau, W.-M.; Luo, Y.; Mattson, M.P.; Tycko, R. Antiparallel β-Sheet Architecture in Iowa-Mutant β-Amyloid Fibrils. Proc. Natl. Acad. Sci. USA 2012, 109, 4443–4448. [Google Scholar] [CrossRef] [PubMed]
  89. Gu, L.; Tran, J.; Jiang, L.; Guo, Z. A New Structural Model of Alzheimer’s Aβ42 Fibrils Based on Electron Paramagnetic Resonance Data and Rosetta Modeling. J. Struct. Biol. 2016, 194, 61–67. [Google Scholar] [CrossRef] [PubMed]
  90. Paparcone, R.; Pires, M.A.; Buehler, M.J. Mutations Alter the Geometry and Mechanical Properties of Alzheimer’s Aβ(1−40) Amyloid Fibrils. Biochemistry 2010, 49, 8967–8977. [Google Scholar] [CrossRef]
  91. Petkova, A.T.; Yau, W.-M.; Tycko, R. Experimental Constraints on Quaternary Structure in Alzheimer’s β-Amyloid Fibrils. Biochemistry 2006, 45, 498–512. [Google Scholar] [CrossRef] [PubMed]
  92. Paravastu, A.K.; Leapman, R.D.; Yau, W.-M.; Tycko, R. Molecular Structural Basis for Polymorphism in Alzheimer’s β-Amyloid Fibrils. Proc. Natl. Acad. Sci. USA 2008, 105, 18349–18354. [Google Scholar] [CrossRef] [PubMed]
  93. Friedemann, M.; Helk, E.; Tiiman, A.; Zovo, K.; Palumaa, P.; Tõugu, V. Effect of Methionine-35 Oxidation on the Aggregation of Amyloid-β Peptide. Biochem. Biophys. Rep. 2015, 3, 94–99. [Google Scholar] [CrossRef] [PubMed]
  94. Bertini, I.; Gonnelli, L.; Luchinat, C.; Mao, J.; Nesi, A. A New Structural Model of Aβ40 Fibrils. J. Am. Chem. Soc. 2011, 133, 16013–16022. [Google Scholar] [CrossRef]
  95. Paravastu, A.K.; Petkova, A.T.; Tycko, R. Polymorphic Fibril Formation by Residues 10–40 of the Alzheimer’s β-Amyloid Peptide. Biophys. J. 2006, 90, 4618–4629. [Google Scholar] [CrossRef] [PubMed]
  96. Petkova, A.T.; Ishii, Y.; Balbach, J.J.; Antzutkin, O.N.; Leapman, R.D.; Delaglio, F.; Tycko, R. A Structural Model for Alzheimer’s β-Amyloid Fibrils Based on Experimental Constraints from Solid State NMR. Proc. Natl. Acad. Sci. USA 2002, 99, 16742–16747. [Google Scholar] [CrossRef] [PubMed]
  97. Karplus, M. Molecular Dynamics Simulations of Biomolecules. Acc. Chem. Res. 2002, 35, 321–323. [Google Scholar] [CrossRef]
  98. Karplus, M.; McCammon, J.A. Molecular Dynamics Simulations of Biomolecules. Nat. Struct. Biol. 2002, 9, 646–652. [Google Scholar] [CrossRef] [PubMed]
  99. Maier, J.A.; Martinez, C.; Kasavajhala, K.; Wickstrom, L.; Hauser, K.E.; Simmerling, C. ff14SB: Improving the Accuracy of Protein Side Chain and Backbone Parameters from ff99SB. J. Chem. Theory Comput. 2015, 11, 3696–3713. [Google Scholar] [CrossRef] [PubMed]
  100. Schmid, N.; Eichenberger, A.P.; Choutko, A.; Riniker, S.; Winger, M.; Mark, A.E.; Van Gunsteren, W.F. Definition and Testing of the GROMOS Force-Field Versions 54A7 and 54B7. Eur. Biophys. J. 2011, 40, 843–856. [Google Scholar] [CrossRef]
  101. Mackerell, A.D.; Feig, M.; Brooks, C.L. Extending the Treatment of Backbone Energetics in Protein Force Fields: Limitations of Gas-phase Quantum Mechanics in Reproducing Protein Conformational Distributions in Molecular Dynamics Simulations. J. Comput. Chem. 2004, 25, 1400–1415. [Google Scholar] [CrossRef]
  102. Harder, E.; Damm, W.; Maple, J.; Wu, C.; Reboul, M.; Xiang, J.Y.; Wang, L.; Lupyan, D.; Dahlgren, M.K.; Knight, J.L.; et al. OPLS3: A Force Field Providing Broad Coverage of Drug-like Small Molecules and Proteins. J. Chem. Theory Comput. 2016, 12, 281–296. [Google Scholar] [CrossRef] [PubMed]
  103. Song, D.; Luo, R.; Chen, H.-F. The IDP-Specific Force Field ff14IDPSFF Improves the Conformer Sampling of Intrinsically Disordered Proteins. J. Chem. Inf. Model. 2017, 57, 1166–1178. [Google Scholar] [CrossRef] [PubMed]
  104. Song, D.; Wang, W.; Ye, W.; Ji, D.; Luo, R.; Chen, H. ff14IDPs Force Field Improving the Conformation Sampling of Intrinsically Disordered Proteins. Chem. Biol. Drug Des. 2017, 89, 5–15. [Google Scholar] [CrossRef] [PubMed]
  105. Zhang, Y.; Liu, H.; Yang, S.; Luo, R.; Chen, H.-F. Well-Balanced Force Field Ff 03 CMAP for Folded and Disordered Proteins. J. Chem. Theory Comput. 2019, 15, 6769–6780. [Google Scholar] [CrossRef] [PubMed]
  106. Huang, J.; Rauscher, S.; Nawrocki, G.; Ran, T.; Feig, M.; De Groot, B.L.; Grubmüller, H.; MacKerell, A.D. CHARMM36m: An Improved Force Field for Folded and Intrinsically Disordered Proteins. Nat. Methods 2017, 14, 71–73. [Google Scholar] [CrossRef]
  107. Liu, H.; Song, D.; Zhang, Y.; Yang, S.; Luo, R.; Chen, H.-F. Extensive Tests and Evaluation of the CHARMM36IDPSFF Force Field for Intrinsically Disordered Proteins and Folded Proteins. Phys. Chem. Chem. Phys. 2019, 21, 21918–21931. [Google Scholar] [CrossRef] [PubMed]
  108. Robustelli, P.; Piana, S.; Shaw, D.E. Developing a Molecular Dynamics Force Field for Both Folded and Disordered Protein States. Proc. Natl. Acad. Sci. USA 2018, 115, E4758–E4766. [Google Scholar] [CrossRef] [PubMed]
  109. Ye, W.; Ji, D.; Wang, W.; Luo, R.; Chen, H.-F. Test and Evaluation of ff99IDPs Force Field for Intrinsically Disordered Proteins. J. Chem. Inf. Model. 2015, 55, 1021–1029. [Google Scholar] [CrossRef]
  110. Best, R.B.; Mittal, J. Protein Simulations with an Optimized Water Model: Cooperative Helix Formation and Temperature-Induced Unfolded State Collapse. J. Phys. Chem. B 2010, 114, 14916–14923. [Google Scholar] [CrossRef]
  111. Piana, S.; Lindorff-Larsen, K.; Shaw, D.E. How Robust Are Protein Folding Simulations with Respect to Force Field Parameterization? Biophys. J. 2011, 100, L47–L49. [Google Scholar] [CrossRef] [PubMed]
  112. Samantray, S.; Yin, F.; Kav, B.; Strodel, B. Different Force Fields Give Rise to Different Amyloid Aggregation Pathways in Molecular Dynamics Simulations. J. Chem. Inf. Model. 2020, 60, 6462–6475. [Google Scholar] [CrossRef]
  113. Rahman, M.U.; Rehman, A.U.; Liu, H.; Chen, H.-F. Comparison and Evaluation of Force Fields for Intrinsically Disordered Proteins. J. Chem. Inf. Model. 2020, 60, 4912–4923. [Google Scholar] [CrossRef] [PubMed]
  114. Shabane, P.S.; Izadi, S.; Onufriev, A.V. General Purpose Water Model Can Improve Atomistic Simulations of Intrinsically Disordered Proteins. J. Chem. Theory Comput. 2019, 15, 2620–2634. [Google Scholar] [CrossRef] [PubMed]
  115. Amber24. Available online: https://ambermd.org/AmberModels_proteins.php (accessed on 19 February 2025).
  116. Carballo-Pacheco, M.; Strodel, B. Comparison of Force Fields for Alzheimer’s A: A Case Study for Intrinsically Disordered Proteins. Protein Sci. 2017, 26, 174–185. [Google Scholar] [CrossRef] [PubMed]
  117. Man, V.H.; He, X.; Derreumaux, P.; Ji, B.; Xie, X.-Q.; Nguyen, P.H.; Wang, J. Effects of All-Atom Molecular Mechanics Force Fields on Amyloid Peptide Assembly: The Case of Aβ16–22 Dimer. J. Chem. Theory Comput. 2019, 15, 1440–1452. [Google Scholar] [CrossRef] [PubMed]
  118. Krupa, P.; Quoc Huy, P.D.; Li, M.S. Properties of Monomeric Aβ42 Probed by Different Sampling Methods and Force Fields: Role of Energy Components. J. Chem. Phys. 2019, 151, 055101. [Google Scholar] [CrossRef]
  119. Carballo-Pacheco, M.; Ismail, A.E.; Strodel, B. On the Applicability of Force Fields To Study the Aggregation of Amyloidogenic Peptides Using Molecular Dynamics Simulations. J. Chem. Theory Comput. 2018, 14, 6063–6075. [Google Scholar] [CrossRef] [PubMed]
  120. Minary, P.; Tuckerman, M.E.; Martyna, G.J. Long Time Molecular Dynamics for Enhanced Conformational Sampling in Biomolecular Systems. Phys. Rev. Lett. 2004, 93, 150201. [Google Scholar] [CrossRef]
  121. Lei, H.; Duan, Y. Improved Sampling Methods for Molecular Simulation. Curr. Opin. Struct. Biol. 2007, 17, 187–191. [Google Scholar] [CrossRef] [PubMed]
  122. Darve, E.; Rodríguez-Gómez, D.; Pohorille, A. Adaptive Biasing Force Method for Scalar and Vector Free Energy Calculations. J. Chem. Phys. 2008, 128, 144120. [Google Scholar] [CrossRef]
  123. Perez, D.; Uberuaga, B.P.; Shim, Y.; Amar, J.G.; Voter, A.F. Chapter 4 Accelerated Molecular Dynamics Methods: Introduction and Recent Developments. In Annual Reports in Computational Chemistry; Elsevier: Amsterdam, The Netherlands, 2009; Volume 5, pp. 79–98. ISBN 978-0-444-53359-3. [Google Scholar]
  124. Zuckerman, D.M. Equilibrium Sampling in Biomolecular Simulations. Annu. Rev. Biophys. 2011, 40, 41–62. [Google Scholar] [CrossRef] [PubMed]
  125. Naqvi, A.A.T.; Mohammad, T.; Hasan, G.M.; Hassan, M.I. Advancements in Docking and Molecular Dynamics Simulations Towards Ligand-Receptor Interactions and Structure-Function Relationships. CTMC 2018, 18, 1755–1768. [Google Scholar] [CrossRef]
  126. Dror, R.O.; Dirks, R.M.; Grossman, J.P.; Xu, H.; Shaw, D.E. Biomolecular Simulation: A Computational Microscope for Molecular Biology. Annu. Rev. Biophys. 2012, 41, 429–452. [Google Scholar] [CrossRef] [PubMed]
  127. Deuflhard, P.; Hermans, J.; Leimkuhler, B.; Mark, A.E.; Reich, S.; Skeel, R.D. (Eds.) Computational Molecular Dynamics: Challenges, Methods, Ideas: Proceedings of the 2nd International Symposium on Algorithms for Macromolecular Modelling, Berlin, May 21–24, 1997; Lecture Notes in Computational Science and Engineering; Springer: Berlin/Heidelberg, Germany, 1999; Volume 4, ISBN 978-3-540-63242-9. [Google Scholar]
  128. Tuckerman, M.E.; Martyna, G.J. Understanding Modern Molecular Dynamics: Techniques and Applications. J. Phys. Chem. B 2000, 104, 159–178. [Google Scholar] [CrossRef]
  129. Becker, O.M.; MacKerell, A.D., Jr.; Roux, B.; Watanabe, M. Chapter 3: Dynamics Methods. In Computational Biochemistry and Biophysics; Becker, O.M., Watanabe, M., Eds.; Marcel Dekker, Inc.: Basel, Switzerland, 2001; ISBN 0-8247-0455-X. [Google Scholar]
  130. Rauscher, S.; Neale, C.; Pomès, R. Simulated Tempering Distributed Replica Sampling, Virtual Replica Exchange, and Other Generalized-Ensemble Methods for Conformational Sampling. J. Chem. Theory Comput. 2009, 5, 2640–2662. [Google Scholar] [CrossRef] [PubMed]
  131. Gelpi, J.; Hospital, A.; Goñi, R.; Orozco, M. Molecular Dynamics Simulations: Advances and Applications. AABC 2015, 37–47. [Google Scholar] [CrossRef]
  132. Huang, X.; Bowman, G.R.; Pande, V.S. Convergence of Folding Free Energy Landscapes via Application of Enhanced Sampling Methods in a Distributed Computing Environment. J. Chem. Phys. 2008, 128, 205106. [Google Scholar] [CrossRef] [PubMed]
  133. Hansmann, U.H.E. Parallel Tempering Algorithm for Conformational Studies of Biological Molecules. Chem. Phys. Lett. 1997, 281, 140–150. [Google Scholar] [CrossRef]
  134. Hansmann, U.H.E.; Okamoto, Y. New Monte Carlo Algorithms for Protein Folding. Curr. Opin. Struct. Biol. 1999, 9, 177–183. [Google Scholar] [CrossRef]
  135. Nagasima, T.; Sugita, Y.; Mitsutake, A.; Okamoto, Y. Generalized-Ensemble Simulations of Spin Systems and Protein Systems. Comput. Phys. Commun. 2002, 146, 69–76. [Google Scholar] [CrossRef]
  136. Lyubartsev, A.P.; Martsinovski, A.A.; Shevkunov, S.V.; Vorontsov-Velyaminov, P.N. New Approach to Monte Carlo Calculation of the Free Energy: Method of Expanded Ensembles. J. Chem. Phys. 1992, 96, 1776–1783. [Google Scholar] [CrossRef]
  137. Marinari, E.; Parisi, G. Simulated Tempering: A New Monte Carlo Scheme. Europhys. Lett. 1992, 19, 451–458. [Google Scholar] [CrossRef]
  138. Swendsen, R.H.; Wang, J.-S. Replica Monte Carlo Simulation of Spin-Glasses. Phys. Rev. Lett. 1986, 57, 2607–2609. [Google Scholar] [CrossRef]
  139. Geyer, C.J. Markov Chain Monte Carlo Maximum Likelihood. In Proceedings of the 23rd Symposium on the Interface, Washington, DC, USA, 21–24 April 1991; pp. 156–163. [Google Scholar]
  140. Geyer, C.J.; Thompson, E.A. Annealing Markov Chain Monte Carlo with Applications to Ancestral Inference. J. Am. Stat. Assoc. 1995, 90, 909–920. [Google Scholar] [CrossRef]
  141. Tesi, M.C.; Janse Van Rensburg, E.J.; Orlandini, E.; Whittington, S.G. Monte Carlo Study of the Interacting Self-Avoiding Walk Model in Three Dimensions. J. Stat. Phys. 1996, 82, 155–181. [Google Scholar] [CrossRef]
  142. Hukushima, K.; Nemoto, K. Exchange Monte Carlo Method and Application to Spin Glass Simulations. J. Phys. Soc. Jpn. 1996, 65, 1604–1608. [Google Scholar] [CrossRef]
  143. Marinari, E.; Parisi, G.; Ruiz-Lorenzo, J.J. Numerical Simulations of Spin Glass Systems. In Series on Directions in Condensed Matter Physics; World Scientific: Singapore, 1997; Volume 12, pp. 59–98. ISBN 978-981-02-3183-5. [Google Scholar]
  144. Sugita, Y.; Okamoto, Y. Replica-Exchange Molecular Dynamics Method for Protein Folding. Chem. Phys. Lett. 1999, 314, 141–151. [Google Scholar] [CrossRef]
  145. Opps, S.B.; Schofield, J. Extended State-Space Monte Carlo Methods. Phys. Rev. E 2001, 63, 056701. [Google Scholar] [CrossRef]
  146. Kofke, D.A. On the Acceptance Probability of Replica-Exchange Monte Carlo Trials. J. Chem. Phys. 2002, 117, 6911–6914. [Google Scholar] [CrossRef]
  147. Rhee, Y.M.; Pande, V.S. Multiplexed-Replica Exchange Molecular Dynamics Method for Protein Folding Simulation. Biophys. J. 2003, 84, 775–786. [Google Scholar] [CrossRef] [PubMed]
  148. Paschek, D.; García, A.E. Reversible Temperature and Pressure Denaturation of a Protein Fragment: A Replica Exchange Molecular Dynamics Simulation Study. Phys. Rev. Lett. 2004, 93, 238105. [Google Scholar] [CrossRef] [PubMed]
  149. Rathore, N.; Chopra, M.; De Pablo, J.J. Optimal Allocation of Replicas in Parallel Tempering Simulations. J. Chem. Phys. 2005, 122, 024111. [Google Scholar] [CrossRef] [PubMed]
  150. Liu, P.; Kim, B.; Friesner, R.A.; Berne, B.J. Replica Exchange with Solute Tempering: A Method for Sampling Biological Systems in Explicit Water. Proc. Natl. Acad. Sci. USA 2005, 102, 13749–13754. [Google Scholar] [CrossRef]
  151. Li, H.; Li, G.; Berg, B.A.; Yang, W. Finite Reservoir Replica Exchange to Enhance Canonical Sampling in Rugged Energy Surfaces. J. Chem. Phys. 2006, 125, 144902. [Google Scholar] [CrossRef]
  152. Trebst, S.; Troyer, M.; Hansmann, U.H.E. Optimized Parallel Tempering Simulations of Proteins. J. Chem. Phys. 2006, 124, 174903. [Google Scholar] [CrossRef] [PubMed]
  153. Brenner, P.; Sweet, C.R.; VonHandorf, D.; Izaguirre, J.A. Accelerating the Replica Exchange Method through an Efficient All-Pairs Exchange. J. Chem. Phys. 2007, 126, 074103. [Google Scholar] [CrossRef] [PubMed]
  154. Okur, A.; Roe, D.R.; Cui, G.; Hornak, V.; Simmerling, C. Improving Convergence of Replica-Exchange Simulations through Coupling to a High-Temperature Structure Reservoir. J. Chem. Theory Comput. 2007, 3, 557–568. [Google Scholar] [CrossRef] [PubMed]
  155. Rick, S.W. Replica Exchange with Dynamical Scaling. J. Chem. Phys. 2007, 126, 054102. [Google Scholar] [CrossRef]
  156. Ruscio, J.Z.; Fawzi, N.L.; Head-Gordon, T. How Hot? Systematic Convergence of the Replica Exchange Method Using Multiple Reservoirs. J. Comput. Chem. 2010, 31, 620–627. [Google Scholar] [CrossRef] [PubMed]
  157. Rauscher, S.; Pomès, R. Simulated Tempering Distributed Replica Sampling: A Practical Guide to Enhanced Conformational Sampling. J. Phys.Conf. Ser. 2010, 256, 012011. [Google Scholar] [CrossRef]
  158. Chen, J.; Im, W.; Brooks, C.L. Application of Torsion Angle Molecular Dynamics for Efficient Sampling of Protein Conformations. J. Comput. Chem. 2005, 26, 1565–1578. [Google Scholar] [CrossRef] [PubMed]
  159. Yun, M.; Lavery, R.; Mousseau, N.; Zakrzewska, K.; Derreumaux, P. ARTIST: An Activated Method in Internal Coordinate Space for Sampling Protein Energy Landscapes. Proteins 2006, 63, 967–975. [Google Scholar] [CrossRef]
  160. Trylska, J.; Tozzini, V.; McCammon, J.A. Exploring Global Motions and Correlations in the Ribosome. Biophys. J. 2005, 89, 1455–1463. [Google Scholar] [CrossRef] [PubMed]
  161. Monticelli, L.; Kandasamy, S.K.; Periole, X.; Larson, R.G.; Tieleman, D.P.; Marrink, S.-J. The MARTINI Coarse-Grained Force Field: Extension to Proteins. J. Chem. Theory Comput. 2008, 4, 819–834. [Google Scholar] [CrossRef] [PubMed]
  162. Chen, J.; Brooks, C.L.; Khandogin, J. Recent Advances in Implicit Solvent-Based Methods for Biomolecular Simulations. Curr. Opin. Struct. Biol. 2008, 18, 140–148. [Google Scholar] [CrossRef] [PubMed]
  163. Grubmüller, H. Predicting Slow Structural Transitions in Macromolecular Systems: Conformational Flooding. Phys. Rev. E 1995, 52, 2893–2906. [Google Scholar] [CrossRef] [PubMed]
  164. Lange, O.F.; Schäfer, L.V.; Grubmüller, H. Flooding in GROMACS: Accelerated Barrier Crossings in Molecular Dynamics. J. Comput. Chem. 2006, 27, 1693–1702. [Google Scholar] [CrossRef] [PubMed]
  165. Huber, T.; Torda, A.E.; Van Gunsteren, W.F. Local Elevation: A Method for Improving the Searching Properties of Molecular Dynamics Simulation. J. Comput.-Aided Mol. Des. 1994, 8, 695–708. [Google Scholar] [CrossRef] [PubMed]
  166. Laio, A.; Parrinello, M. Escaping Free-Energy Minima. Proc. Natl. Acad. Sci. USA 2002, 99, 12562–12566. [Google Scholar] [CrossRef]
  167. Bussi, G.; Laio, A.; Parrinello, M. Equilibrium Free Energies from Nonequilibrium Metadynamics. Phys. Rev. Lett. 2006, 96, 090601. [Google Scholar] [CrossRef] [PubMed]
  168. Leone, V.; Marinelli, F.; Carloni, P.; Parrinello, M. Targeting Biomolecular Flexibility with Metadynamics. Curr. Opin. Struct. Biol. 2010, 20, 148–154. [Google Scholar] [CrossRef] [PubMed]
  169. Darve, E.; Pohorille, A. Calculating Free Energies Using Average Force. J. Chem. Phys. 2001, 115, 9169–9183. [Google Scholar] [CrossRef]
  170. Torrie, G.M.; Valleau, J.P. Nonphysical Sampling Distributions in Monte Carlo Free-Energy Estimation: Umbrella Sampling. J. Comput. Phys. 1977, 23, 187–199. [Google Scholar] [CrossRef]
  171. Beutler, T.C.; Van Gunsteren, W.F. The Computation of a Potential of Mean Force: Choice of the Biasing Potential in the Umbrella Sampling Technique. J. Chem. Phys. 1994, 100, 1492–1497. [Google Scholar] [CrossRef]
  172. Voter, A.F. A Method for Accelerating the Molecular Dynamics Simulation of Infrequent Events. J. Chem. Phys. 1997, 106, 4665–4677. [Google Scholar] [CrossRef]
  173. Voter, A.F. Hyperdynamics: Accelerated Molecular Dynamics of Infrequent Events. Phys. Rev. Lett. 1997, 78, 3908–3911. [Google Scholar] [CrossRef]
  174. Hamelberg, D.; Mongan, J.; McCammon, J.A. Accelerated Molecular Dynamics: A Promising and Efficient Simulation Method for Biomolecules. J. Chem. Phys. 2004, 120, 11919–11929. [Google Scholar] [CrossRef] [PubMed]
  175. Shen, T.; Hamelberg, D. A Statistical Analysis of the Precision of Reweighting-Based Simulations. J. Chem. Phys. 2008, 129, 034103. [Google Scholar] [CrossRef] [PubMed]
  176. Fajer, M.; Hamelberg, D.; McCammon, J.A. Replica-Exchange Accelerated Molecular Dynamics (REXAMD) Applied to Thermodynamic Integration. J. Chem. Theory Comput. 2008, 4, 1565–1569. [Google Scholar] [CrossRef] [PubMed]
  177. Wereszczynski, J.; McCammon, J.A. Using Selectively Applied Accelerated Molecular Dynamics to Enhance Free Energy Calculations. J. Chem. Theory Comput. 2010, 6, 3285–3292. [Google Scholar] [CrossRef] [PubMed]
  178. Sinko, W.; De Oliveira, C.A.F.; Pierce, L.C.T.; McCammon, J.A. Protecting High Energy Barriers: A New Equation to Regulate Boost Energy in Accelerated Molecular Dynamics Simulations. J. Chem. Theory Comput. 2012, 8, 17–23. [Google Scholar] [CrossRef]
  179. Wang, Y.; Harrison, C.B.; Schulten, K.; McCammon, J.A. Implementation of Accelerated Molecular Dynamics in NAMD. Comput. Sci. Disc. 2011, 4, 015002. [Google Scholar] [CrossRef] [PubMed]
  180. Pawnikar, S.; Bhattarai, A.; Wang, J.; Miao, Y. Binding Analysis Using Accelerated Molecular Dynamics Simulations and Future Perspectives. AABC 2022, 15, 1–19. [Google Scholar] [CrossRef] [PubMed]
  181. Bernardi, R.C.; Melo, M.C.R.; Schulten, K. Enhanced Sampling Techniques in Molecular Dynamics Simulations of Biological Systems. Biochim. Biophys. Acta (BBA)-Gen. Subj. 2015, 1850, 872–877. [Google Scholar] [CrossRef] [PubMed]
  182. Harpole, T.J.; Delemotte, L. Conformational Landscapes of Membrane Proteins Delineated by Enhanced Sampling Molecular Dynamics Simulations. Biochim. Biophys. Acta (BBA) Biomembr. 2018, 1860, 909–926. [Google Scholar] [CrossRef] [PubMed]
  183. Smirnov, G.S.; Stegailov, V.V. Efficiency of Classical Molecular Dynamics Algorithms on Supercomputers. Math. Models Comput. Simul. 2016, 8, 734–743. [Google Scholar] [CrossRef]
  184. Jung, J.; Kobayashi, C.; Kasahara, K.; Tan, C.; Kuroda, A.; Minami, K.; Ishiduki, S.; Nishiki, T.; Inoue, H.; Ishikawa, Y.; et al. New Parallel Computing Algorithm of Molecular Dynamics for Extremely Huge Scale Biological Systems. J. Comput. Chem. 2021, 42, 231–241. [Google Scholar] [CrossRef] [PubMed]
  185. Guo, Z.; Lu, D.; Yan, Y.; Hu, S.; Liu, R.; Tan, G.; Sun, N.; Jiang, W.; Liu, L.; Chen, Y.; et al. Extending the Limit of Molecular Dynamics with Ab Initio Accuracy to 10 Billion Atoms. In Proceedings of the 27th ACM SIGPLAN Symposium on Principles and Practice of Parallel Programming, Seoul, Republic of Korea, 2–6 April 2022; ACM: New York, NY, USA, 2022; pp. 205–218. [Google Scholar]
  186. Shaw, D.E.; Adams, P.J.; Azaria, A.; Bank, J.A.; Batson, B.; Bell, A.; Bergdorf, M.; Bhatt, J.; Butts, J.A.; Correia, T.; et al. Anton 3: Twenty Microseconds of Molecular Dynamics Simulation before Lunch. In Proceedings of the International Conference for High Performance Computing, Networking, Storage and Analysis. St. Louis, MO, USA, 14–19 November 2021; ACM: New York, NY, USA, 2021; pp. 1–11. [Google Scholar]
  187. Top500 The List. Available online: https://top500.org/lists/top500/2023/06/ (accessed on 20 June 2023).
  188. Lumi. Available online: https://www.lumi-supercomputer.eu/efficient-molecular-dynamics-simulations-on-lumi/ (accessed on 5 February 2025).
  189. Shirts, M.; Pande, V.S. Screen Savers of the World Unite! Science 2000, 290, 1903–1904. [Google Scholar] [CrossRef] [PubMed]
  190. Folding@home. Available online: http://foldingathome.org (accessed on 6 February 2025).
  191. Zimmerman, M.I.; Porter, J.R.; Ward, M.D.; Singh, S.; Vithani, N.; Meller, A.; Mallimadugula, U.L.; Kuhn, C.E.; Borowsky, J.H.; Wiewiora, R.P.; et al. SARS-CoV-2 Simulations Go Exascale to Predict Dramatic Spike Opening and Cryptic Pockets across the Proteome. Nat. Chem. 2021, 13, 651–659. [Google Scholar] [CrossRef]
  192. Lemkul, J.A.; Bevan, D.R. The Role of Molecular Simulations in the Development of Inhibitors of Amyloid β-Peptide Aggregation for the Treatment of Alzheimer’s Disease. ACS Chem. Neurosci. 2012, 3, 845–856. [Google Scholar] [CrossRef]
  193. Nasica-Labouze, J.; Nguyen, P.H.; Sterpone, F.; Berthoumieu, O.; Buchete, N.-V.; Coté, S.; De Simone, A.; Doig, A.J.; Faller, P.; Garcia, A.; et al. Amyloid β Protein and Alzheimer’s Disease: When Computer Simulations Complement Experimental Studies. Chem. Rev. 2015, 115, 3518–3563. [Google Scholar] [CrossRef] [PubMed]
  194. Rodríguez, M.H.; Morales, L.G.F.; Basurto, J.C.; Hernández, M.C.R. Molecular Docking and Molecular Dynamics Simulation to Evaluate Compounds That Avoid the Amyloid Beta 1-42 Aggregation. In Computational Modeling of Drugs Against Alzheimer’s Disease; Roy, K., Ed.; Neuromethods; Springer: New York, NY, USA, 2018; Volume 132, pp. 229–248. ISBN 978-1-4939-7403-0. [Google Scholar]
  195. Low, K.J.Y.; Venkatraman, A.; Mehta, J.S.; Pervushin, K. Molecular Mechanisms of Amyloid Disaggregation. J. Adv. Res. 2022, 36, 113–132. [Google Scholar] [CrossRef] [PubMed]
  196. Wise, R.A. Dopamine, Learning and Motivation. Nat. Rev. Neurosci. 2004, 5, 483–494. [Google Scholar] [CrossRef]
  197. Berridge, C.W.; Waterhouse, B.D. The Locus Coeruleus–Noradrenergic System: Modulation of Behavioral State and State-Dependent Cognitive Processes. Brain Res. Rev. 2003, 42, 33–84. [Google Scholar] [CrossRef]
  198. Murchison, C.F.; Zhang, X.-Y.; Zhang, W.-P.; Ouyang, M.; Lee, A.; Thomas, S.A. A Distinct Role for Norepinephrine in Memory Retrieval. Cell 2004, 117, 131–143. [Google Scholar] [CrossRef] [PubMed]
  199. Ono, K.; Hasegawa, K.; Naiki, H.; Yamada, M. Anti-Parkinsonian Agents Have Anti-Amyloidogenic Activity for Alzheimer’s β-Amyloid Fibrils in Vitro. Neurochem. Int. 2006, 48, 275–285. [Google Scholar] [CrossRef] [PubMed]
  200. Huong, V.T.; Shimanouchi, T.; Shimauchi, N.; Yagi, H.; Umakoshi, H.; Goto, Y.; Kuboi, R. Catechol Derivatives Inhibit the Fibril Formation of Amyloid-β Peptides. J. Biosci. Bioeng. 2010, 109, 629–634. [Google Scholar] [CrossRef] [PubMed]
  201. Tavanti, F.; Pedone, A.; Menziani, M.C. Computational Insight into the Effect of Natural Compounds on the Destabilization of Preformed Amyloid-β(1–40) Fibrils. Molecules 2018, 23, 1320. [Google Scholar] [CrossRef]
  202. Chen, Y.; Li, X.; Zhan, C.; Lao, Z.; Li, F.; Dong, X.; Wei, G. A Comprehensive Insight into the Mechanisms of Dopamine in Disrupting Aβ Protofibrils and Inhibiting Aβ Aggregation. ACS Chem. Neurosci. 2021, 12, 4007–4019. [Google Scholar] [CrossRef]
  203. Zou, Y.; Qian, Z.; Chen, Y.; Qian, H.; Wei, G.; Zhang, Q. Norepinephrine Inhibits Alzheimer’s Amyloid-β Peptide Aggregation and Destabilizes Amyloid-β Protofibrils: A Molecular Dynamics Simulation Study. ACS Chem. Neurosci. 2019, 10, 1585–1594. [Google Scholar] [CrossRef]
  204. Gao, D.; Wan, J.; Zou, Y.; Gong, Y.; Dong, X.; Xu, Z.; Tang, J.; Wei, G.; Zhang, Q. The Destructive Mechanism of Aβ1–42 Protofibrils by Norepinephrine Revealed via Molecular Dynamics Simulations. Phys. Chem. Chem. Phys. 2022, 24, 19827–19836. [Google Scholar] [CrossRef] [PubMed]
  205. Gong, Y.; Zhan, C.; Zou, Y.; Qian, Z.; Wei, G.; Zhang, Q. Serotonin and Melatonin Show Different Modes of Action on Aβ42 Protofibril Destabilization. ACS Chem. Neurosci. 2021, 12, 799–809. [Google Scholar] [CrossRef]
  206. Pal, S.; Paul, S. ATP Controls the Aggregation of Aβ16–22 Peptides. J. Phys. Chem. B 2020, 124, 210–223. [Google Scholar] [CrossRef] [PubMed]
  207. Rodríguez, J.J.; Noristani, H.N.; Verkhratsky, A. The Serotonergic System in Ageing and Alzheimer’s Disease. Prog. Neurobiol. 2012, 99, 15–41. [Google Scholar] [CrossRef]
  208. Porter, R.J.; Lunn, B.S.; O’Brien, J.T. Effects of Acute Tryptophan Depletion on Cognitive Function in Alzheimer’s Disease and in the Healthy Elderly. Psychol. Med. 2003, 33, 41–49. [Google Scholar] [CrossRef] [PubMed]
  209. Hardy, J.A.; Wester, P.; Backstrom, I.; Gottfries, J.; Oreland, L.; Stenstrom, A.; Winblad, B. The Regional Distribution of Dopamine and Serotonin Uptake and Transmitter Concentrations in the Human Brain. Neurochem. Int. 1987, 10, 445–450. [Google Scholar] [CrossRef]
  210. Hornedo-Ortega, R.; Da Costa, G.; Cerezo, A.B.; Troncoso, A.M.; Richard, T.; Garcia-Parrilla, M.C. In Vitro Effects of Serotonin, Melatonin, and Other Related Indole Compounds on Amyloid-β Kinetics and Neuroprotection. Mol. Nutr. Food Res. 2018, 62, 1700383. [Google Scholar] [CrossRef] [PubMed]
  211. Karasek, M. Melatonin, Human Aging, and Age-Related Diseases. Exp. Gerontol. 2004, 39, 1723–1729. [Google Scholar] [CrossRef]
  212. Sack, R.L.; Lewy, A.J.; Erb, D.L.; Vollmer, W.M.; Singer, C.M. Human Melatonin Production Decreases with Age. J. Pineal Res. 1986, 3, 379–388. [Google Scholar] [CrossRef]
  213. Skene, D. Melatonin Rhythmicity: Effect of Age and Alzheimer’s Disease. Exp. Gerontol. 2003, 38, 199–206. [Google Scholar] [CrossRef]
  214. Skene, D.J.; Vivien-Roels, B.; Sparks, D.L.; Hunsaker, J.C.; Pévet, P.; Ravid, D.; Swaab, D.F. Daily Variation in the Concentration of Melatonin and 5-Methoxytryptophol in the Human Pineal Gland: Effect of Age and Alzheimer’s Disease. Brain Res. 1990, 528, 170–174. [Google Scholar] [CrossRef] [PubMed]
  215. Lin, L.; Huang, Q.-X.; Yang, S.-S.; Chu, J.; Wang, J.-Z.; Tian, Q. Melatonin in Alzheimer’s Disease. Int. J. Mol. Sci. 2013, 14, 14575–14593. [Google Scholar] [CrossRef]
  216. Zhang, S.; Wang, P.; Ren, L.; Hu, C.; Bi, J. Protective Effect of Melatonin on Soluble Aβ1–42-Induced Memory Impairment, Astrogliosis, and Synaptic Dysfunction via the Musashi1/Notch1/Hes1 Signaling Pathway in the Rat Hippocampus. Alzheimer’s Res. Ther. 2016, 8, 40. [Google Scholar] [CrossRef]
  217. Poeggeler, B.; Miravalle, L.; Zagorski, M.G.; Wisniewski, T.; Chyan, Y.-J.; Zhang, Y.; Shao, H.; Bryant-Thomas, T.; Vidal, R.; Frangione, B.; et al. Melatonin Reverses the Profibrillogenic Activity of Apolipoprotein E4 on the Alzheimer Amyloid Aβ Peptide. Biochemistry 2001, 40, 14995–15001. [Google Scholar] [CrossRef] [PubMed]
  218. Pappolla, M.; Bozner, P.; Soto, C.; Shao, H.; Robakis, N.K.; Zagorski, M.; Frangione, B.; Ghiso, J. Inhibition of Alzheimer β-Fibrillogenesis by Melatonin. J. Biol. Chem. 1998, 273, 7185–7188. [Google Scholar] [CrossRef] [PubMed]
  219. Patel, A.; Malinovska, L.; Saha, S.; Wang, J.; Alberti, S.; Krishnan, Y.; Hyman, A.A. ATP as a Biological Hydrotrope. Science 2017, 356, 753–756. [Google Scholar] [CrossRef]
  220. Kuramochi, M.; Nakamura, M.; Takahashi, H.; Komoriya, T.; Takita, T.; Pham, N.T.K.; Yasukawa, K.; Yoshimune, K. Adenosine Triphosphate Induces Amorphous Aggregation of Amyloid β by Increasing Aβ Dynamics. Sci. Rep. 2024, 14, 8134. [Google Scholar] [CrossRef] [PubMed]
  221. Coskuner, O.; Murray, I.V.J. Adenosine Triphosphate (ATP) Reduces Amyloid-β Protein Misfolding in Vitro. JAD 2014, 41, 561–574. [Google Scholar] [CrossRef]
  222. Ngo, S.T.; Fang, S.-T.; Huang, S.-H.; Chou, C.-L.; Huy, P.D.Q.; Li, M.S.; Chen, Y.-C. Anti-Arrhythmic Medication Propafenone a Potential Drug for Alzheimer’s Disease Inhibiting Aggregation of Aβ: In Silico and in Vitro Studies. J. Chem. Inf. Model. 2016, 56, 1344–1356. [Google Scholar] [CrossRef] [PubMed]
  223. Juszczak, G.R.; Swiergiel, A.H. Properties of Gap Junction Blockers and Their Behavioural, Cognitive and Electrophysiological Effects: Animal and Human Studies. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2009, 33, 181–198. [Google Scholar] [CrossRef]
  224. Rozental, R.; Srinivas, M.; Spray, D.C. How to Close a Gap Junction Channel: Efficacies and Potencies of Uncoupling Agents. In Connexin Methods and Protocols; Humana Press: Totowa, NJ, USA, 2000; Volume 154, pp. 447–476. ISBN 978-1-59259-043-8. [Google Scholar]
  225. Hellmich, H.L.; Rojo, D.R.; Micci, M.-A.; Sell, S.L.; Boone, D.R.; Crookshanks, J.M.; DeWitt, D.S.; Masel, B.E.; Prough, D.S. Pathway Analysis Reveals Common Pro-Survival Mechanisms of Metyrapone and Carbenoxolone after Traumatic Brain Injury. PLoS ONE 2013, 8, e53230. [Google Scholar] [CrossRef] [PubMed]
  226. Thakur, P.; Nehru, B. Inhibition of Neuroinflammation and Mitochondrial Dysfunctions by Carbenoxolone in the Rotenone Model of Parkinson’s Disease. Mol. Neurobiol. 2015, 51, 209–219. [Google Scholar] [CrossRef]
  227. Sandeep, T.C.; Yau, J.L.W.; MacLullich, A.M.J.; Noble, J.; Deary, I.J.; Walker, B.R.; Seckl, J.R. 11β-Hydroxysteroid Dehydrogenase Inhibition Improves Cognitive Function in Healthy Elderly Men and Type 2 Diabetics. Proc. Natl. Acad. Sci. USA 2004, 101, 6734–6739. [Google Scholar] [CrossRef] [PubMed]
  228. Sharma, S.; Nehru, B.; Saini, A. Inhibition of Alzheimer’s Amyloid-Beta Aggregation in-Vitro by Carbenoxolone: Insight into Mechanism of Action. Neurochem. Int. 2017, 108, 481–493. [Google Scholar] [CrossRef] [PubMed]
  229. Gianni, L.; Bellotti, V.; Gianni, A.M.; Merlini, G. New Drug Therapy of Amyloidoses: Resorption of AL-Type Deposits with 4′-Iodo-4′-Deoxydoxorubicin. Blood 1995, 86, 855–861. [Google Scholar] [CrossRef] [PubMed]
  230. Merlini, G.; Ascari, E.; Amboldi, N.; Bellotti, V.; Arbustini, E.; Perfetti, V.; Ferrari, M.; Zorzoli, I.; Marinone, M.G.; Garini, P. Interaction of the Anthracycline 4′-Iodo-4′-Deoxydoxorubicin with Amyloid Fibrils: Inhibition of Amyloidogenesis. Proc. Natl. Acad. Sci. USA 1995, 92, 2959–2963. [Google Scholar] [CrossRef]
  231. Palha, J.A.; Ballinari, D.; Amboldi, N.; Cardoso, I.; Fernandes, R.; Bellotti, V.; Merlini, G.; Saraiva, M.J. 4′-Iodo-4′-Deoxydoxorubicin Disrupts the Fibrillar Structure of Transthyretin Amyloid. Am. J. Pathol. 2000, 156, 1919–1925. [Google Scholar] [CrossRef]
  232. Tagliavini, F.; McArthur, R.A.; Canciani, B.; Giaccone, G.; Porro, M.; Bugiani, M.; Lievens, P.M.-J.; Bugiani, O.; Peri, E.; Dall’Ara, P.; et al. Effectiveness of Anthracycline Against Experimental Prion Disease in Syrian Hamsters. Science 1997, 276, 1119–1121. [Google Scholar] [CrossRef] [PubMed]
  233. Diomede, L.; Cassata, G.; Fiordaliso, F.; Salio, M.; Ami, D.; Natalello, A.; Doglia, S.M.; De Luigi, A.; Salmona, M. Tetracycline and Its Analogues Protect Caenorhabditis Elegans from β Amyloid-Induced Toxicity by Targeting Oligomers. Neurobiol. Dis. 2010, 40, 424–431. [Google Scholar] [CrossRef] [PubMed]
  234. Balducci, C.; Santamaria, G.; La Vitola, P.; Brandi, E.; Grandi, F.; Viscomi, A.R.; Beeg, M.; Gobbi, M.; Salmona, M.; Ottonello, S.; et al. Doxycycline Counteracts Neuroinflammation Restoring Memory in Alzheimer’s Disease Mouse Models. Neurobiol. Aging 2018, 70, 128–139. [Google Scholar] [CrossRef]
  235. Forloni, G.; Colombo, L.; Girola, L.; Tagliavini, F.; Salmona, M. Anti-amyloidogenic Activity of Tetracyclines: Studies in Vitro. FEBS Lett. 2001, 487, 404–407. [Google Scholar] [CrossRef] [PubMed]
  236. Cardoso, I.; Merlini, G.; Saraiva, M.J. 4 ′-iodo-4′-Deoxydoxorubicin and Tetracyclines Disrupt Transthyretin Amyloid Fibrils in Vitro Producing Noncytotoxic Species: Screening for TTR Fibril Disrupters. FASEB J. 2003, 17, 803–809. [Google Scholar] [CrossRef]
  237. Cardoso, I.; Saraiva, M.J. Doxycycline Disrupts Transthyretin Amyloid: Evidence from Studies in a FAP Transgenic Mice Model. FASEB J. 2006, 20, 234–239. [Google Scholar] [CrossRef] [PubMed]
  238. Obici, L.; Cortese, A.; Lozza, A.; Lucchetti, J.; Gobbi, M.; Palladini, G.; Perlini, S.; Saraiva, M.J.; Merlini, G. Doxycycline plus Tauroursodeoxycholic Acid for Transthyretin Amyloidosis: A Phase II Study. Amyloid 2012, 19, 34–36. [Google Scholar] [CrossRef] [PubMed]
  239. Stoilova, T.; Colombo, L.; Forloni, G.; Tagliavini, F.; Salmona, M. A New Face for Old Antibiotics: Tetracyclines in Treatment of Amyloidoses. J. Med. Chem. 2013, 56, 5987–6006. [Google Scholar] [CrossRef]
  240. Airoldi, C.; Colombo, L.; Manzoni, C.; Sironi, E.; Natalello, A.; Doglia, S.M.; Forloni, G.; Tagliavini, F.; Del Favero, E.; Cantù, L.; et al. Tetracycline Prevents Aβ Oligomer Toxicity through an Atypical Supramolecular Interaction. Org. Biomol. Chem. 2011, 9, 463–472. [Google Scholar] [CrossRef]
  241. Montagna, G.; Cazzulani, B.; Obici, L.; Uggetti, C.; Giorgetti, S.; Porcari, R.; Ruggiero, R.; Mangione, P.P.; Brambilla, M.; Lucchetti, J.; et al. Benefit of Doxycycline Treatment on Articular Disability Caused by Dialysis Related Amyloidosis. Amyloid 2013, 20, 173–178. [Google Scholar] [CrossRef] [PubMed]
  242. Giorgetti, S.; Raimondi, S.; Pagano, K.; Relini, A.; Bucciantini, M.; Corazza, A.; Fogolari, F.; Codutti, L.; Salmona, M.; Mangione, P.; et al. Effect of Tetracyclines on the Dynamics of Formation and Destructuration of β2-Microglobulin Amyloid Fibrils. J. Biol. Chem. 2011, 286, 2121–2131. [Google Scholar] [CrossRef] [PubMed]
  243. Ward, J.E.; Ren, R.; Toraldo, G.; SooHoo, P.; Guan, J.; O’Hara, C.; Jasuja, R.; Trinkaus-Randall, V.; Liao, R.; Connors, L.H.; et al. Doxycycline Reduces Fibril Formation in a Transgenic Mouse Model of AL Amyloidosis. Blood 2011, 118, 6610–6617. [Google Scholar] [CrossRef]
  244. Gautieri, A.; Beeg, M.; Gobbi, M.; Rigoldi, F.; Colombo, L.; Salmona, M. The Anti-Amyloidogenic Action of Doxycycline: A Molecular Dynamics Study on the Interaction with Aβ42. Int. J. Mol. Sci. 2019, 20, 4641. [Google Scholar] [CrossRef]
  245. Kamal, M.; Knox, J.; Horne, R.I.; Tiwari, O.S.; Burns, A.R.; Han, D.; Levy, D.; Laor Bar-Yosef, D.; Gazit, E.; Vendruscolo, M.; et al. A Rapid in Vivo Pipeline to Identify Small Molecule Inhibitors of Amyloid Aggregation. Nat. Commun. 2024, 15, 8311. [Google Scholar] [CrossRef] [PubMed]
  246. Vasilev, B.; Atanasova, M. A (Comprehensive) Review of the Application of Quantitative Structure–Activity Relationship (QSAR) in the Prediction of New Compounds with Anti-Breast Cancer Activity. Appl. Sci. 2025, 15, 1206. [Google Scholar] [CrossRef]
Figure 1. The factors contributing to elevated Aβ levels in both forms of Alzheimer’s disease (AD) are illustrated on the left. Mutations in the APP gene, as well as in the PSEN1 and PSEN2 genes, which encode the catalytic subunits (presenilins 1 and 2) of γ-secretase, inheritance of the APOE ε4 allele, increased activity of β-secretase (BACE1), and impaired Aβ clearance all contribute to Aβ accumulation, reinforcing the amyloid cascade hypothesis The APP cleavage pathways are shown schematically. In the non-amyloidogenic pathway, the cleavage by α-secretase and γ-secretase generates soluble APPα (sAPPα), p3 (composed of 17–40/42 amino acids), and the C-terminal fragment α (CTFα, C83), containing 83 amino acids. In the amyloidogenic pathway, β-secretase first cleaves APP, resulting in sAPPβ, and subsequently, γ-secretase generates Aβ (1–40/42 amino acids) and CTFβ (C99), which consists of 99 amino acids. The increased levels of Aβ can be cleared, degraded, or undergo oligomerization. Based on Aβ’s secondary structure, three potential self-aggregation pathways via primary nucleation have been identified: specific on-pathways, off-pathways, and a non-specific amorphous pathway. Amyloid fibrils are formed from the primary nucleation of U-shaped β-strand (β-hairpin) monomers, which then aggregate into a nucleus, low- and high-molecular-weight fibrillar oligomers, and amyloid protofibrils (pathway 1). Alternatively, an irregular β-sheet monomer self-associates through an irregular nucleus and oligomers, eventually forming non-toxic aggregates through the off-pathway (pathway 2). The third pathway leads to the creation of amorphous, unstructured aggregates (tangles) composed of disordered random coil monomers. Secondary mechanisms for aggregate formation, depending on fibril concentration, are also outlined, including fibril fragmentation (monomer-independent), elongation, and secondary nucleation (monomer-dependent).
Figure 1. The factors contributing to elevated Aβ levels in both forms of Alzheimer’s disease (AD) are illustrated on the left. Mutations in the APP gene, as well as in the PSEN1 and PSEN2 genes, which encode the catalytic subunits (presenilins 1 and 2) of γ-secretase, inheritance of the APOE ε4 allele, increased activity of β-secretase (BACE1), and impaired Aβ clearance all contribute to Aβ accumulation, reinforcing the amyloid cascade hypothesis The APP cleavage pathways are shown schematically. In the non-amyloidogenic pathway, the cleavage by α-secretase and γ-secretase generates soluble APPα (sAPPα), p3 (composed of 17–40/42 amino acids), and the C-terminal fragment α (CTFα, C83), containing 83 amino acids. In the amyloidogenic pathway, β-secretase first cleaves APP, resulting in sAPPβ, and subsequently, γ-secretase generates Aβ (1–40/42 amino acids) and CTFβ (C99), which consists of 99 amino acids. The increased levels of Aβ can be cleared, degraded, or undergo oligomerization. Based on Aβ’s secondary structure, three potential self-aggregation pathways via primary nucleation have been identified: specific on-pathways, off-pathways, and a non-specific amorphous pathway. Amyloid fibrils are formed from the primary nucleation of U-shaped β-strand (β-hairpin) monomers, which then aggregate into a nucleus, low- and high-molecular-weight fibrillar oligomers, and amyloid protofibrils (pathway 1). Alternatively, an irregular β-sheet monomer self-associates through an irregular nucleus and oligomers, eventually forming non-toxic aggregates through the off-pathway (pathway 2). The third pathway leads to the creation of amorphous, unstructured aggregates (tangles) composed of disordered random coil monomers. Secondary mechanisms for aggregate formation, depending on fibril concentration, are also outlined, including fibril fragmentation (monomer-independent), elongation, and secondary nucleation (monomer-dependent).
Pharmaceuticals 18 00306 g001
Figure 2. The primary and secondary structures of human Aβ1–42 the known aggregate shapes, as retrieved from the Protein Data Bank (www.rcsb.org), are depicted. The peptide chains are color-coded according to specific regions in the primary amino acid sequence: the N-terminus (hydrophilic or metal-binding region) from D1 to Q15 is colored beige; the central hydrophobic core (CHC) or β1 region from K16 to A21 is shown in cyan; the loop or central hydrophilic region from E22 to K28 is yellow; the second hydrophobic region (β2) from G29 to M35 is colored salmon; and the C-terminal region from V36 to A42 is green. The PDB codes are labeled above the corresponding structures. A possible mechanism for oligomer formation is the rapid assembly driven by hydrophobic interactions, including those involving the C-terminus [38].
Figure 2. The primary and secondary structures of human Aβ1–42 the known aggregate shapes, as retrieved from the Protein Data Bank (www.rcsb.org), are depicted. The peptide chains are color-coded according to specific regions in the primary amino acid sequence: the N-terminus (hydrophilic or metal-binding region) from D1 to Q15 is colored beige; the central hydrophobic core (CHC) or β1 region from K16 to A21 is shown in cyan; the loop or central hydrophilic region from E22 to K28 is yellow; the second hydrophobic region (β2) from G29 to M35 is colored salmon; and the C-terminal region from V36 to A42 is green. The PDB codes are labeled above the corresponding structures. A possible mechanism for oligomer formation is the rapid assembly driven by hydrophobic interactions, including those involving the C-terminus [38].
Pharmaceuticals 18 00306 g002
Figure 3. Depiction of the endogenous compounds studied through MD simulations.
Figure 3. Depiction of the endogenous compounds studied through MD simulations.
Pharmaceuticals 18 00306 g003
Figure 4. The structures of the repurposed drugs (Propafenone (PPF), Carbenoxolone (CBX), and Doxycycline (DXC) studied through MD simulations are presented.
Figure 4. The structures of the repurposed drugs (Propafenone (PPF), Carbenoxolone (CBX), and Doxycycline (DXC) studied through MD simulations are presented.
Pharmaceuticals 18 00306 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Atanasova, M. Small-Molecule Inhibitors of Amyloid Beta: Insights from Molecular Dynamics—Part A: Endogenous Compounds and Repurposed Drugs. Pharmaceuticals 2025, 18, 306. https://doi.org/10.3390/ph18030306

AMA Style

Atanasova M. Small-Molecule Inhibitors of Amyloid Beta: Insights from Molecular Dynamics—Part A: Endogenous Compounds and Repurposed Drugs. Pharmaceuticals. 2025; 18(3):306. https://doi.org/10.3390/ph18030306

Chicago/Turabian Style

Atanasova, Mariyana. 2025. "Small-Molecule Inhibitors of Amyloid Beta: Insights from Molecular Dynamics—Part A: Endogenous Compounds and Repurposed Drugs" Pharmaceuticals 18, no. 3: 306. https://doi.org/10.3390/ph18030306

APA Style

Atanasova, M. (2025). Small-Molecule Inhibitors of Amyloid Beta: Insights from Molecular Dynamics—Part A: Endogenous Compounds and Repurposed Drugs. Pharmaceuticals, 18(3), 306. https://doi.org/10.3390/ph18030306

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop